0% found this document useful (0 votes)
250 views987 pages

Technical Thermodynamics For Engineers: Achim Schmidt

Uploaded by

Victor Palacios
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
250 views987 pages

Technical Thermodynamics For Engineers: Achim Schmidt

Uploaded by

Victor Palacios
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 987

Achim 

Schmidt

Technical
Thermodynamics
for Engineers
Basics and Applications
Second Edition
Technical Thermodynamics for Engineers
Achim Schmidt

Technical Thermodynamics
for Engineers
Basics and Applications
Second Edition
Achim Schmidt
Fakultät Technik und Informatik
Department Maschinenbau
University of Applied Sciences, Hamburg
Hamburg, Germany

ISBN 978-3-030-97149-6 ISBN 978-3-030-97150-2 (eBook)


https://doi.org/10.1007/978-3-030-97150-2

© Springer Nature Switzerland AG 2019, 2022


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Dedicated to my parents.
Preface

The following quote is attributed to the famous physicist Arnold Sommerfeld and is
probably the beginning of many thermodynamics textbooks:
Thermodynamics is a funny subject. The first time you go through it, you do not understand
it at all. The second time you go through it, you think you understand it, except for one or
two small points. The third time you go through it, you know you do not understand it, but
by that time you are so used to it, it does not bother you anymore.

In fact, thermodynamics is probably one of the most difficult and challenging subjects
of mechanical engineering studies. Many students claim that it would be a great
subject—if only it weren’t for the written exam.
Now that I have given the lectures in Technical Thermodynamics I/II for several
semesters at the University of Applied Sciences in Hamburg, I have decided, to write
an accompanying textbook. I believe that my personal gaps of understanding have
become smaller. According to Arnold Sommerfeld, I’m one of those who’ve dealt
with thermodynamics for at least the third time. Interestingly enough, my fascination
for the subject is still growing. In addition, each semester there are a couple of
questions from the students that cause me to constantly question the theory and that
give me a new perspective on the subject. I have collected these and my questions
during my studies and tried to summarise the answers as well as possible in this
textbook.
Why another textbook—there are already so many available! During the prepara-
tion of my lectures, I had the impression that there are many great textbooks on the
subject, each with very specific merits. Nevertheless, my approach is to combine what
I find to be best understandable and to go in the depths where it is required, always in
order not to lose the red, thermodynamic thread. Numerous questions of the students
have been helpful, especially in the context of the exam preparations. So this book is
a reference guide that has the same structure as my lecture. Although English is not
my mother tongue, it was very important to me to write the book in English: Firstly,
I reach a larger readership than if the book were in German. Secondly, linguistic
limitations force me to explain even complicated things in a simple way.

vii
viii Preface

This book is categorised in three parts: Part I introduces the fundamentals of


technical thermodynamics. The first and second laws of thermodynamics are derived,
that both enable us to understand the principle of energy conversion. Fluids are
simply treated as ideal gases or incompressible liquids. The physical description of
these fluids obeys equations of state. Thermodynamic cycles are discussed, which
convert thermal energy into mechanical energy, e.g. an internal combustion engine.
Furthermore, thermodynamic cycles can be utilised to shift thermal energy from a low
temperature level to a larger temperature level, e.g. in a fridge. In Part II, real fluids
are investigated, i.e. fluids that can change their aggregate state, for example. These
fluids cannot be treated as easily as the fluids from Part I. Furthermore, mixtures
of fluids are introduced, e.g. humid air as a mixture of dry air and water. Changes
of state of these mixtures are treated as well. However, these mixtures will not
be chemically reactive. Finally, Part III includes chemically reactive fluids. This
is required to calculate combustion processes, for instance. Combustion processes
are part of many technical applications and require the knowledge of the basics
introduced before.
I have tried to find a good mixture of theory, examples and tasks. Since technical
drawings are the language of the engineer, this book contains numerous detailed
illustrations which are intended to clarify even the most difficult aspects of theory.
In the second edition, the fluid properties and state diagrams in the appendix have
been thoroughly revised and supplemented. In addition, an exemplary formulary has
been added, which may be useful for exams. Further tasks are added. Finally, minor
corrections and improvements have been made.
On this occasion, I would like to express my sincere thanks to the readers of
the first edition, whose feedback provided me with important contributions for the
revision.
Finally, any of your feedback is highly appreciated!

Hamburg, Germany Achim Schmidt


Summer 2021
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 How Is This Book Structured? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Classification of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.1 Technical Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Statistical Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.3 Chemical Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Distinction Thermodynamics/Heat Transfer . . . . . . . . . . . . . . . . . . 7
1.3.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.2 Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4 History of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.1 The Caloric Theory Around 1780 . . . . . . . . . . . . . . . . . . . 8
1.4.2 Thermodynamics as from the 18th Century . . . . . . . . . . . 9
1.4.3 Thermodynamics in the 21st Century . . . . . . . . . . . . . . . . 13
1.4.4 Modern Automotive Applications . . . . . . . . . . . . . . . . . . . 18

Part I Basics and Ideal Fluids


2 Energy and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1 Mechanical Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.1 Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1.2 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.1.3 Spring Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2 Thermal Energy—Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3 Chemical Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 Changeability of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4.1 Joule’s Paddle Wheel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4.2 Internal Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3 System and State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1 System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.1 Classification of Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 37

ix
x Contents

3.1.2 Permeability of Systems—Open Versus Closed


Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.3 Examples for Thermodynamic Systems . . . . . . . . . . . . . . 43
3.2 State of a System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2.1 Thermal State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.2 Caloric State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.2.3 Outer State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2.4 Size of a System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2.5 Extensive, Intensive and Specific State Values . . . . . . . . . 49
4 Thermodynamic Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.1 Mechanical Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.2 Thermal Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3 Chemical Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.4 Local Thermodynamic Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.5 Assumptions in Technical Thermodynamics . . . . . . . . . . . . . . . . . . 60
5 Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.1 Gibbs’ Phase Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.1.1 Single-Component Systems Without Phase
Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.1.2 Single-Component Systems with Phase Change . . . . . . . 70
5.1.3 Multi-Component Systems . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.2 Explicit Versus Implicit Equations of State . . . . . . . . . . . . . . . . . . . 72
6 Thermal Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.1 Temperature Variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.2 Pressure Variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.3 Ideal Gas Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7 Changes of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
7.1 The p, v-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.1.1 Isothermal Change of State . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.1.2 Isobaric Change of State . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
7.1.3 Isochoric Change of State . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.2 Equilibrium Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7.2.1 Quasi-Static Changes of State . . . . . . . . . . . . . . . . . . . . . . 91
7.2.2 Requirement for a Quasi-Static Change of State . . . . . . . 93
7.3 Reversible Versus Irreversible Changes of State . . . . . . . . . . . . . . . 94
7.3.1 Mechanical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.3.2 Thermal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
7.3.3 Chemical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7.4 Conventional Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Contents xi

8 Thermodynamic Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111


8.1 Equilibrium Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
8.2 Transient State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
8.3 Thermodynamic Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
8.4 Steady State Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.4.1 Open Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.4.2 Closed Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
8.4.3 Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
9 Process Values Heat and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
9.1 Thermal Energy—Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
9.2 Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
9.2.1 Definition of Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
9.2.2 Volume Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
9.2.3 Effective Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
9.2.4 Systems with Internal Friction—Dissipation . . . . . . . . . . 133
9.2.5 Dissipation Versus Outer Friction . . . . . . . . . . . . . . . . . . . 138
9.2.6 Mechanical Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
9.2.7 Shaft Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
9.2.8 Shifting Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
9.2.9 Technical Work Respectively Pressure Work . . . . . . . . . . 144
10 State Value Versus Process Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
10.1 Total Differential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
10.2 Schwarz’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
11 First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
11.1 Principle of Equivalence Between Work and Heat . . . . . . . . . . . . . 157
11.2 Closed Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
11.2.1 Systems at Rest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
11.2.2 Systems in Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
11.2.3 Partial Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
11.3 Open Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
11.3.1 Formulation of the First Law of Thermodynamics
for Open Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
11.3.2 Non-steady State Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
11.3.3 Steady State Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
11.3.4 Partial Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
12 Caloric Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
12.1 Specific Internal Energy u and Specific Enthalpy h
for Ideal Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
12.2 Specific Entropy s as New State Value for Ideal Gases . . . . . . . . . 217
12.3 Derivation of the Caloric Equations for Real Fluids . . . . . . . . . . . 222
12.3.1 Specific Internal Energy u . . . . . . . . . . . . . . . . . . . . . . . . . . 222
12.3.2 Specific Enthalpy h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
12.4 Handling of the Caloric State Equations . . . . . . . . . . . . . . . . . . . . . 232
xii Contents

12.4.1 Ideal Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233


12.4.2 Distinction Between cv and c p for Ideal Gases . . . . . . . . 233
12.4.3 Isentropic Exponent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
12.4.4 Temperature Dependent Specific Heat Capacity . . . . . . . 239
12.4.5 Incompressible Fluids, Solids . . . . . . . . . . . . . . . . . . . . . . . 244
12.4.6 Adiabatic Throttle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
13 Meaning and Handling of Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
13.1 Entropy—Clarification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
13.2 Comparison Entropy Balance Versus First Law
of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
13.3 Energy Conversion—Why Do We Need Entropy? . . . . . . . . . . . . . 258
13.4 The T, s-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
13.4.1 Benefit of a New State Diagram . . . . . . . . . . . . . . . . . . . . . 262
13.4.2 Physical Laws in a T, s-Diagram for Ideal Gases . . . . . . 264
13.5 Adiabatic, Reversible Change of State . . . . . . . . . . . . . . . . . . . . . . . 267
13.6 Polytropic Change of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
13.7 Entropy Balancing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
13.7.1 Entropy Balance for Closed Systems . . . . . . . . . . . . . . . . 281
13.7.2 Entropy Balance for Open Systems . . . . . . . . . . . . . . . . . . 285
13.7.3 Thermodynamic Mean Temperature . . . . . . . . . . . . . . . . . 289
13.7.4 Entropy and Process Evaluation . . . . . . . . . . . . . . . . . . . . . 294
13.8 Entropy—Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
14 Transient Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
14.1 Mechanical Driven Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
14.2 Thermal Driven Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
14.3 Chemical Driven Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
14.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
15 Second Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
15.1 Formulation According to Planck—Clockwise Cycle
Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
15.1.1 The Thermal Engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
15.1.2 Why Clockwise Cycle? . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
15.2 Formulation According to Clausius—Counterclockwise
Cycle processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
15.2.1 The Cooling Machine/Heat Pump . . . . . . . . . . . . . . . . . . . 364
15.2.2 Why Counterclockwise Cycle? . . . . . . . . . . . . . . . . . . . . . 368
15.3 The Carnot-Machine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
15.3.1 The Carnot-Machine—Clockwise Cycle . . . . . . . . . . . . . 372
15.3.2 The Carnot-Machine—Counterclockwise Cycle . . . . . . . 375
Contents xiii

16 Exergy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
16.1 Exergy of Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
16.1.1 Heat at Constant Temperature . . . . . . . . . . . . . . . . . . . . . . 384
16.1.2 Heat at Variable Temperature . . . . . . . . . . . . . . . . . . . . . . . 385
16.1.3 Sign of the Exergy of Heat . . . . . . . . . . . . . . . . . . . . . . . . . 388
16.2 Exergy of Fluid Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
16.3 Exergy of Closed Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
16.4 Loss of Exergy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
16.4.1 Closed System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
16.4.2 Open System in Steady State Operation . . . . . . . . . . . . . . 401
16.4.3 Thermodynamic Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
16.5 Sankey-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
16.5.1 Open System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
16.5.2 Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413
17 Components and Thermodynamic Cycles . . . . . . . . . . . . . . . . . . . . . . . 423
17.1 Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
17.1.1 Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
17.1.2 Compressor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
17.1.3 Thermal Turbomachines in a h, s-Diagram . . . . . . . . . . . 431
17.1.4 Adiabatic Throttle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
17.1.5 Heat Exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
17.2 Thermodynamic Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
17.2.1 Carnot Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
17.2.2 Joule Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
17.2.3 Clausius Rankine Process . . . . . . . . . . . . . . . . . . . . . . . . . . 445
17.2.4 Seiliger Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
17.2.5 Stirling Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
17.2.6 Compression Heat Pump . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
17.2.7 Process Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454

Part II Real Fluids and Mixtures


18 Single-Component Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
18.1 Ideal Gas Versus Real Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
18.2 Phase Change Real Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476
18.2.1 Example: Isobaric Vaporisation . . . . . . . . . . . . . . . . . . . . . 476
18.2.2 The p, v, T-state Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
18.2.3 p, T -Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
18.2.4 T, v-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486
18.2.5 p, v-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487
18.2.6 State Description Within the Wet Steam Region . . . . . . . 490
18.3 State Values of Real Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
18.3.1 Van der Waals Equation of State . . . . . . . . . . . . . . . . . . . . 493
18.3.2 Redlich-Kwong . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
18.3.3 Peng-Robinson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
xiv Contents

18.3.4 Berthelot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498


18.3.5 Dieterici . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
18.3.6 Virial Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
18.3.7 Steam Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
18.4 Energetic Consideration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
18.4.1 Reversibility of Vaporisation . . . . . . . . . . . . . . . . . . . . . . . 505
18.4.2 Heat of Vaporisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
18.4.3 Caloric State Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510
18.4.4 Clausius-Clapeyron Relation . . . . . . . . . . . . . . . . . . . . . . . 518
18.5 Adiabatic Throttling—Joule-Thomson Effect . . . . . . . . . . . . . . . . . 521
18.5.1 Ideal Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523
18.5.2 Real Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
19 Mixture of Ideal Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537
19.1 Concentration Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 538
19.2 Dalton’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
19.3 Laws of Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545
19.3.1 Concentration, Thermal State Values . . . . . . . . . . . . . . . . 545
19.3.2 Internal Energy, Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . 548
19.3.3 Adiabatic Mixing Temperature . . . . . . . . . . . . . . . . . . . . . . 550
19.3.4 Irreversibility of Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
20 Humid Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565
20.1 Thermodynamic State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
20.1.1 Concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
20.1.2 Aggregate State of the Water . . . . . . . . . . . . . . . . . . . . . . . 567
20.1.3 Distinction Between Vaporisation and Evaporation . . . . 569
20.1.4 Unsaturated Versus Saturated Air . . . . . . . . . . . . . . . . . . . 570
20.2 Specific State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
20.2.1 Thermal State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
20.2.2 Caloric State Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
20.2.3 Specific Enthalpy h 1+x . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
20.2.4 Specific Entropy s1+x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 588
20.2.5 Overview Possible Cases . . . . . . . . . . . . . . . . . . . . . . . . . . 592
20.3 The h 1+x , x-diagram According to Mollier . . . . . . . . . . . . . . . . . . . 602
20.4 Changes of State for Humid Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604
20.4.1 Heating and Cooling at Constant Water Content . . . . . . . 605
20.4.2 Dehumidification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607
20.4.3 Adiabatic Mixing of Humid Air . . . . . . . . . . . . . . . . . . . . . 610
20.4.4 Humidification of Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
20.4.5 Adiabatic Saturation Temperature . . . . . . . . . . . . . . . . . . . 617
20.4.6 The h 1+x , x-Diagram for Varying Total Pressure . . . . . . 621
Contents xv

21 Steady State Flow Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 631


21.1 Incompressible Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
21.2 Adiabatic Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
21.2.1 Adiabatic Diffusor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
21.2.2 Adiabatic Nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 639
21.3 Velocity of Sound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
21.4 Fanno Correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 648
21.5 Rayleigh Correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659
21.6 Normal Shock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
21.7 Supersonic Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 670
21.7.1 Flow of a Converging Nozzle . . . . . . . . . . . . . . . . . . . . . . . 670
21.7.2 Laval-Nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 678
22 Thermodynamic Cycles with Phase Change . . . . . . . . . . . . . . . . . . . . . . 689
22.1 Steam Power Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
22.1.1 Clausius–Rankine Process . . . . . . . . . . . . . . . . . . . . . . . . . 690
22.1.2 Steam Power Plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
22.2 Heat Pump and Cooling Machine . . . . . . . . . . . . . . . . . . . . . . . . . . . 701
22.2.1 Mechanical Compression . . . . . . . . . . . . . . . . . . . . . . . . . . 701
22.2.2 Thermal Compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705

Part III Reactive Systems


23 Combustion Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
23.1 Fossil Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
23.2 Fuel Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 722
23.2.1 Solid Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 722
23.2.2 Liquid Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
23.2.3 Gaseous Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724
23.3 Stoichiometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724
23.3.1 Solid/Liquid Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725
23.3.2 Gaseous Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 732
23.3.3 Mass Conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740
23.3.4 Conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740
23.3.5 Setting Up a Chemical Equation . . . . . . . . . . . . . . . . . . . . 743
23.3.6 Dew Point of the Exhaust Gas . . . . . . . . . . . . . . . . . . . . . . 746
23.4 Energetic Balancing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 757
23.4.1 Lower Heating Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 758
23.4.2 Conceptual 3-Steps Combustion . . . . . . . . . . . . . . . . . . . . 760
23.4.3 Higher Heating Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 771
23.4.4 Combustion Calculation Component
by Component . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 775
23.4.5 Molar and Volume Specific Lower/Higher Heating
Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 777
23.4.6 Combustion Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 779
23.5 Combustion Chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784
xvi Contents

23.5.1 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784


23.5.2 Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785
24 Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 799
24.1 Mass Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 799
24.2 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 804
24.2.1 Caloric Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . 805
24.2.2 Open Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 810
24.2.3 Closed Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 815
24.3 Gibbs Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821
24.3.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821
24.3.2 Molar Gibbs Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 822
24.3.3 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 823
24.4 Chemical Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825
24.4.1 Multi-Component Systems . . . . . . . . . . . . . . . . . . . . . . . . . 825
24.4.2 Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 834
24.5 Exergy of a Fossil Fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 837

Appendix A: Steam Table (Water) According to IAPWS . . . . . . . . . . . . . . 845


Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies . . . . 871
Appendix C: Caloric State Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 949
Appendix D: The h1+x , x-Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 959
Appendix E: Formulary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 963
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 967
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 969
Nomenclature

Roman Symbols

A Area, m2
a Cohesion pressure; see Eq. 18.38, Pa m6 mol–2
a Mass fraction ashes, –
a Speed of sound, m s–1
a Acceleration, m s–2
b Molar co-volume; see Eq. 18.38, m3 mol–1
Ḃx Flux of anergy, W
Bx Anergy, J
B(T ) Virial coefficient, m3 mol–1
bx Specific anergy, J kg–1
C Number of components, –
C Capacity, F
C, cM Molar heat capacity, J mol–1 K–1
C̄ Molar arithmetic mean heat capacity, J mol–1 K–1
C̄¯ Molar logarithmic mean heat capacity, J mol–1 K–1
C Constant
C(T ) Virial coefficient, m6 mol–2
c Mass fraction carbon, −
c Velocity, m s–1
ci Molarity of a component i, mol m–3
c Specific heat capacity, J kg–1 K–1
c̄ Specific, averaged heat capacity, J kg–1 K–1
D(T ) Virial coefficient, m9 mol–3
D, d Diameter, m
E Energy, J
Ex Exergy, J
Ė x Flux of exergy, W
E x,V Loss of exergy, J

xvii
xviii Nomenclature

 Ė x,V Flux of loss of exergy, W


E xm Molar specific absolute exergy, J mol–1
E xm,F Molar specific absolute exergy of a fuel, J mol–1
ex Specific exergy, J kg–1
ex,V Specific loss of exergy, J kg–1
F Degree of freedom, –
F Force, N
F Helmholtz energy, J
f Specific Helmholtz energy, J kg–1
G Gibbs energy, J
Gm Molar specific Gibbs energy, J mol–1
g Specific Gibbs energy, J kg–1
g Gravitational acceleration, g = 9.81 m s–2 , m s–2
H, h Height in a gravity field, m
H Enthalpy, J
Hm Molar specific absolute enthalpy, J mol–1
HU Mass specific lower heating value, J kg–1
HUM Molar specific lower heating value, J mol–1
HUv Volume specific lower heating value, J m–3
H0 Mass specific higher heating value, J kg–1
H0M Molar specific higher heating value, J mol–1
H0v Volume specific higher heating value, J m–3
0B Hm Molar specific enthalpy of formation at standard conditions, J mol–1
0R Hm Molar specific enthalpy of reaction at standard conditions, J mol–1
h Mass fraction hydrogen, –
h Specific enthalpy, J kg–1
h v Specific enthalpy of vaporisation, J kg–1
h m Specific enthalpy of melting, J kg–1


I Momentum, kg m s–1
K Revolutions per working stroke, –
k Heat transition coefficient, W m–2 K–1
kB Boltzmann constant, k B = 1.3806 × 10–23 J K–1 , J K–1
kF Spring constant, N m–1
L min Minimum molar specific air need, –
lmin Minimum mass specific air need, –
M Molar mass, kg mol–1
M Torque, N m
m Mass, kg
ṁ Mass flux, kg s–1
ṁ  Mass flux density, kg s–1 m–2
Ma Mach-number, –
NA Avogadro constant, NA = 6.022045×1023 mol–1 , mol–1
n Mass fraction nitrogen, –
n Molar quantity, mol
Nomenclature xix

n Polytropic exponent, –
n Speed, s–1
ṅ Molar flux, mol s–1

→ n Normal
Omin Minimum molar-specific oxygen need, –
omin Minimum mass-specific oxygen need, –
o Mass fraction oxygen, –
P Number of phases, –
P Power, W
p Pressure, Pa
pi Partial pressure of a component i, Pa
Q Electric charge, F
Q Heat or thermal energy, J
Q̇ Heat flux, W
q Specific heat, J kg–1
R Individual gas constant, J kg–1 K–1
R, r Radius, m
RM General gas constant, RM = 8.3143 J mol–1 K–1 , J mol–1 K–1
S Entropy, J K–1
s Specific entropy, J kg–1 K–1
si Specific entropy generation, J kg–1 K–1
sa Specific entropy carried with heat, J kg–1 K–1
s Mass fraction sulphur, –
s Distance, m
Ṡi Flux of entropy generation, W K–1
Ṡa Flux of entropy carried with heat, W K–1
R Sm Molar specific entropy of a reaction, J mol–1 K–1
Sm Molar specific absolute entropy, J mol–1 K–1
T Absolute thermodynamic temperature, K
Tr Reduced thermodynamic temperature according to Eq. 18.63, –
t Time, s
U Internal energy, J
Um Molar specific absolute internal energy, J mol–1
u Specific internal energy, J kg–1
V Volume, m3
V̇ Volume flux, m3 s–1
V Voltage, V
v Specific volume, m3 kg−1
W Work, J
w Specific work, J kg–1
w Mass fraction water, –
x Vapour ratio, –
x Water content, –
x Molar fraction, –
xi Molar concentration of a component i, –
xx Nomenclature

x Coordinate, m
y Coordinate, m
Y Pressure work, J
y Specific pressure work, J kg–1
Z Compressibility factor, –
Z Extensive state value
z Specific state value, z = mZ
z Coordinate, m
z Distance, m

Greek Symbols

α Abbreviation
α Angle, °
α Heat transfer coefficient, W m–2 K–1
β Isobaric volumetric thermal expansion coefficient, K–1
δh Isenthalpic throttle coefficient, K Pa–1
δT Isothermal throttle coefficient, m3 kg–1
 Compression ratio, –
η Abbreviation
η, ξ Frictional constant, kg s–1
η Efficiency, –
γi Stoichiometric factor of a component i, –
κ Isentropic coefficient, –
λ Air-fuel equivalence ratio, –
μ Chemical potential, J mol–1
μi Mass-specific exhaust gas composition of component i, –
νi Molar-specific exhaust gas composition of component i, –
Ω Flow function of a nozzle, –
Ω Statistical weight, measure of the probability, –
ω Acentric factor, –
Ψ Dissipation, J
Ψ̇ Flux of dissipation, W
ψ Specific dissipation, J kg–1
ψ Specific dissipation per length, J kg–1 m–1
ψ Relative saturation, –
ψi Volume ratio of a component i, –
ρ Density, kg m–3
ρi Partial density of a component i, kg m–3
σi Volume concentration of a component i, –
ε Coefficient of performance, –
ϕ Relative humidity, –
ϑ Celsius-temperature, °C
Nomenclature xxi

ξ Abbreviation
ξi Mass concentration of a component i, –

Subscripts

A Air (wet)
a Air (dry)
a Outer
C Cylinder
comp Compression
cond. Condenser
eff Effective
EG Exhaust gas
el Electric
env Environment
F Fuel
fric. Friction
G Gas
gas Gas
HP High pressure
ice Ice
in Inlet
irrev Irreversible
K Control volume
kin Kinetic
L Liquid
liq Liquid
LP Low pressure
m Mean
m Melting
M,m Molar
max Maximum
Mech Mechanic
min Minimum
MP Medium pressure
n Narrowest cross section
out Outlet
P Piston
p p = const.
pot Potential
R Reservoir
ref Reference state
rev Reversible
xxii Nomenclature

S Steel
s Saturated
shift Shifting
Source Source term
spr Spring
swing Stroke
Sys System
t Technical
T T = const.
th Thermal
total Total
V Volume
v v = const.
V,v Vapour
W Water

Acronyms

 Saturated liquid state, saturated humid air


 Saturated vapour state
δ Process value
d State value
1P Single-phase
2P Two-phase
C Carnot
CM Cold machine/fridge
cp Critical point
HP Heat pump
HP High pressure
HT Heat transfer
HVAC Heating, ventilation and air conditioning technology
LP Low pressure
Pr Product
TE Thermal engine
TP Triple point
List of Figures

Fig. 1.1 What is thermodynamics all about? (Tasks


of thermodynamics) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Fig. 1.2 Distinction thermodynamics (left)/heat transfer (right) . . . . . . . 7
Fig. 1.3 Working principle of a steam machine . . . . . . . . . . . . . . . . . . . . 12
Fig. 1.4 Gas turbine (Schematic working principle) . . . . . . . . . . . . . . . . . 14
Fig. 1.5 Fossil power plant (left: Schematic working principle,
right: Visualisation in a p, v-diagram, liquid water is
supposed approximately to be incompressible) . . . . . . . . . . . . . 15
Fig. 1.6 Cooling machine (left: Schematic working principle,
right: Illustration in a T, s-diagram, the compressor is
supposed to be isentropic) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Fig. 1.7 Thermoelectric generator (TEG) . . . . . . . . . . . . . . . . . . . . . . . . . 18
Fig. 1.8 Waste heat recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Fig. 2.1 a Definition of work and b power . . . . . . . . . . . . . . . . . . . . . . . . 24
Fig. 2.2 Kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Fig. 2.3 Potential energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Fig. 2.4 Spring energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Fig. 2.5 Heat as thermal energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Fig. 2.6 Joule’s paddle wheel experiment . . . . . . . . . . . . . . . . . . . . . . . . . 29
Fig. 2.7 Joule’s paddle wheel experiment, state (1) and (2) . . . . . . . . . . . 30
Fig. 2.8 Sketch to Problem 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Fig. 2.9 Solution to Problem 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Fig. 3.1 Definition of a system, (i) indicates the internal state,
(B) the system boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Fig. 3.2 Example of a bank account . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Fig. 3.3 Heterogeneous versus homogeneous systems . . . . . . . . . . . . . . . 38
Fig. 3.4 Heterogeneous systems—fluidised bed, see [15] . . . . . . . . . . . . 39
Fig. 3.5 Heterogeneous systems—diesel spray formation, see [16] . . . . 40
Fig. 3.6 Aggregate states from left to right: solid/liquid/gas . . . . . . . . . . 40
Fig. 3.7 Permeability of systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

xxiii
xxiv List of Figures

Fig. 3.8 Cooling machine on the left: thermodynamic cycle


on the right: possible system boundaries . . . . . . . . . . . . . . . . . . . 44
Fig. 3.9 Examples for thermodynamic open systems . . . . . . . . . . . . . . . . 45
Fig. 3.10 Overview state values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Fig. 3.11 Overview state values for a closed system . . . . . . . . . . . . . . . . . 47
Fig. 3.12 Extensive and intensive state values . . . . . . . . . . . . . . . . . . . . . . 50
Fig. 3.13 Extensive state value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Fig. 3.14 Intensive versus specific state value according to [1] Left:
thermodynamic equilibrium possible, though ρ1 < ρ2
Right: thermodynamic equilibrium impossible, if p1 = p2 . . . . 52
Fig. 3.15 Definition of molar mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Fig. 4.1 a Imbalanced system, b System in thermodynamic
equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Fig. 4.2 Calculating the pressure in equilibrium state . . . . . . . . . . . . . . . 56
Fig. 4.3 Thermal equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Fig. 4.4 Zeroth law of thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Fig. 4.5 Principle of temperature measurement . . . . . . . . . . . . . . . . . . . . 59
Fig. 4.6 Imbalanced systems in steady state—example of a heat
conducting wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Fig. 4.7 Assumptions in thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . 61
Fig. 4.8 Example thermodynamic equilibrium . . . . . . . . . . . . . . . . . . . . . 61
Fig. 4.9 Sketch to Problem 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Fig. 4.10 Solution to Problem 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Fig. 4.11 Solution to Problem 4.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Fig. 5.1 p, T -diagram for water—Degrees of freedom . . . . . . . . . . . . . . 71
Fig. 5.2 Independent state values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Fig. 6.1 Gay-Lussac law—set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Fig. 6.2 Gay-Lussac law—results for thermodynamic equilibrium . . . . . 77
Fig. 6.3 Boyle-Mariotte law—set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Fig. 6.4 Boyle-Mariotte law—results for thermodynamic
equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Fig. 7.1 Change of state from one equilibrium state to another . . . . . . . . 84
Fig. 7.2 Isothermal change of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Fig. 7.3 Isobaric change of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Fig. 7.4 Isochoric change of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Fig. 7.5 Solution to Problem 7.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Fig. 7.6 Quasi-static versus non-quasi-static change of state:
Illustration in a p,v-diagram for gas compression . . . . . . . . . . . . 91
Fig. 7.7 Path-independent equilibrium state . . . . . . . . . . . . . . . . . . . . . . . 94
Fig. 7.8 Wire pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Fig. 7.9 Dissipation in a closed systems . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Fig. 7.10 Bouncing ball driven by z . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Fig. 7.11 Pressure balancing driven by p . . . . . . . . . . . . . . . . . . . . . . . . . 97
Fig. 7.12 Cooling down of a liquid driven by T . . . . . . . . . . . . . . . . . . . 98
Fig. 7.13 Mixing of gases driven by ξi . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
List of Figures xxv

Fig. 7.14 Overview irreversible/reversible—Example


of an adiabatic compression/expansion . . . . . . . . . . . . . . . . . . . . 100
Fig. 7.15 Analogous model according to [19] . . . . . . . . . . . . . . . . . . . . . . . 102
Fig. 7.16 Approach technical thermodynamics according to [1] . . . . . . . . 103
Fig. 7.17 Isobaric change of state—differential notation . . . . . . . . . . . . . . 104
Fig. 7.18 p, v- and T, s-diagram according to Fig. 7.14 . . . . . . . . . . . . . . 108
Fig. 8.1 Equilibrium process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Fig. 8.2 Transient state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
Fig. 8.3 Thermodynamic cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Fig. 8.4 Stirling cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Fig. 8.5 Steady state flow process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
Fig. 8.6 Steady state closed system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Fig. 8.7 Steady state thermodynamic cycle (black box) . . . . . . . . . . . . . . 117
Fig. 9.1 Thermal equilibrium is achieved by heat flux . . . . . . . . . . . . . . . 120
Fig. 9.2 Sign convention heat—advice . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Fig. 9.3 Definition of work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Fig. 9.4 Sign convention for the process values heat and work . . . . . . . . 125
Fig. 9.5 Deriving volume work—Compression (1) → (2) . . . . . . . . . . . . 126
Fig. 9.6 Specific volume work wv,12 —illustration in a p, v-diagram . . . 126
Fig. 9.7 Effective work Weff —illustration and visualisation
in a p, V -diagram, a Compression b Expansion . . . . . . . . . . . . 130
Fig. 9.8 Dissipation in closed (a) and open systems (b) . . . . . . . . . . . . . 133
Fig. 9.9 Left: compression to the same volume, right: expansion
to the same pressure (both adiabatic) (1) → (2)
reversible, (1) → (2 ) irreversible . . . . . . . . . . . . . . . . . . . . . . . . 135
Fig. 9.10 Dissipation in closed systems due to fluid turbulence
(a), outer friction (b) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Fig. 9.11 Illustration of mechanical work . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Fig. 9.12 Shaft work for a closed system . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Fig. 9.13 Shifting work in open systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Fig. 9.14 Technical work—no dissipation, outer energies ignored . . . . . . 144
Fig. 9.15 Equivalent model for technical work @ pressure decrease,
related to Fig. 9.14b . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Fig. 10.1 Process versus state value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Fig. 10.2 Total differential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Fig. 10.3 Schwarz’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Fig. 11.1 Equivalence of heat and work . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Fig. 11.2 Converting thermal to mechanical energy . . . . . . . . . . . . . . . . . . 159
Fig. 11.3 Closed system at rest—supply of heat . . . . . . . . . . . . . . . . . . . . . 159
Fig. 11.4 Closed system at rest—supply of work . . . . . . . . . . . . . . . . . . . . 160
Fig. 11.5 Closed system at rest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
Fig. 11.6 Converting energy in a closed system . . . . . . . . . . . . . . . . . . . . . 161
Fig. 11.7 Closed system—no cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
Fig. 11.8 Closed system in motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
xxvi List of Figures

Fig. 11.9 First law of thermodynamics for closed systems,


see Sect. 3.2.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
Fig. 11.10 Partial energy equation for closed systems . . . . . . . . . . . . . . . . . 166
Fig. 11.11 Effective work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Fig. 11.12 a Supply of electrical energy in an isochoric system,
b Supply of electrical energy in a non-isochoric system,
c Supply of electrical energy with an electrical capacitor . . . . . 169
Fig. 11.13 a Reversible, adiabatic compression, b Irreversible,
adiabatic compression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
Fig. 11.14 Partial energy equation for closed systems . . . . . . . . . . . . . . . . . 173
Fig. 11.15 Partial energy equation for closed systems . . . . . . . . . . . . . . . . . 174
Fig. 11.16 Solution Problem 11.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
Fig. 11.17 Solution to Problem 11.4—Isobaric change of state . . . . . . . . . . 180
Fig. 11.18 Solution to Problem 11.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
Fig. 11.19 Numerical solution to Problem 11.5 . . . . . . . . . . . . . . . . . . . . . . 186
Fig. 11.20 Compression process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
Fig. 11.21 Schematic diagram (equivalent model) . . . . . . . . . . . . . . . . . . . . 188
Fig. 11.22 From closed to open systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
Fig. 11.23 Energy balance open system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
Fig. 11.24 Energy within an open system . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Fig. 11.25 Mass balance open system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
Fig. 11.26 Energy crossing the system boundary of an open system . . . . . 193
Fig. 11.27 First law of thermodynamics for an open system . . . . . . . . . . . . 195
Fig. 11.28 Sketch to Problem 11.6, see also Fig. 9.15 . . . . . . . . . . . . . . . . . 197
Fig. 11.29 Sketch to Problem 11.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Fig. 11.30 Open systems with multiple inlets/outlets . . . . . . . . . . . . . . . . . . 205
Fig. 11.31 Technical work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Fig. 11.32 What does technical work mean? . . . . . . . . . . . . . . . . . . . . . . . . . 209
Fig. 11.33 Specific pressure work y12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
Fig. 12.1 Joule expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
Fig. 12.2 Isobaric (Case A) versus isochoric (Case B) change of state . . . 235
Fig. 12.3 Solution to Problem 12.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Fig. 12.4 Temperature dependency of the specific heat capacity . . . . . . . . 240
Fig. 12.5 Sketch to Problem 12.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Fig. 12.6 Sketch to Problem 12.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
Fig. 13.1 Mechanism of entropy in a closed systems . . . . . . . . . . . . . . . . . 254
Fig. 13.2 Balance of entropy in a closed system . . . . . . . . . . . . . . . . . . . . . 256
Fig. 13.3 Balance of energy in a closed system . . . . . . . . . . . . . . . . . . . . . 257
Fig. 13.4 Power plant as a block-box . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
Fig. 13.5 T, s-diagram: Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
Fig. 13.6 T, s-diagram: a Reversible and b irreversible . . . . . . . . . . . . . . . 263
Fig. 13.7 T, s-diagram: Isochore and isobar . . . . . . . . . . . . . . . . . . . . . . . . 264
Fig. 13.8 Isobar and isochore in a T, s-diagram (ideal gas) . . . . . . . . . . . . 266
Fig. 13.9 Adiabatic, reversible change of state . . . . . . . . . . . . . . . . . . . . . . 267
List of Figures xxvii

Fig. 13.10 Reversible compression of an ideal gas: Isothermal


versus adiabatic according to Problem 13.2 . . . . . . . . . . . . . . . . 271
Fig. 13.11 Polytropic change of state for an ideal gas in a p,v-diagram . . . 273
Fig. 13.12 Polytropic change of state for an ideal gas in a p,v-
and in a T, s-diagram according to [1] . . . . . . . . . . . . . . . . . . . . 274
Fig. 13.13 Overview ideal gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
Fig. 13.14 Polytropic change of state for an ideal gas - Compression
(closed system), see Problem 13.2 . . . . . . . . . . . . . . . . . . . . . . . . 277
Fig. 13.15 Polytropic change of state for an ideal gas—Expansion
(closed system), see Problem 13.3 . . . . . . . . . . . . . . . . . . . . . . . . 278
Fig. 13.16 Sketch of the compressor, see Problem 13.4 . . . . . . . . . . . . . . . . 279
Fig. 13.17 Energy and entropy balance for a closed system . . . . . . . . . . . . 281
Fig. 13.18 Isothermal irreversible compression of an ideal gas . . . . . . . . . . 283
Fig. 13.19 a Energy and b entropy balance—open system . . . . . . . . . . . . . 285
Fig. 13.20 Entropy balance—steady state flow system, multiple
inlets/outlets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
Fig. 13.21 Thermodynamic mean temperature—steady state flow
system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
Fig. 13.22 Thermodynamic mean temperature—closed system . . . . . . . . . 292
Fig. 13.23 Lifting a crate of beer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
Fig. 13.24 Sketch to Problem 13.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
Fig. 13.25 Left: Compression to the same volume, right: Expansion
to the same pressure (both adiabatic) . . . . . . . . . . . . . . . . . . . . . . 299
Fig. 13.26 Sketch to Problem 13.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
Fig. 13.27 System boundaries for Problem 13.8 . . . . . . . . . . . . . . . . . . . . . . 301
Fig. 13.28 Massless boundary, Problem 13.8(e) . . . . . . . . . . . . . . . . . . . . . . 304
Fig. 13.29 Sketch to Problem 13.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
Fig. 13.30 Sketch to Problem 13.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
Fig. 13.31 Sketch to Problem 13.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
Fig. 13.32 Sketch to Problem 13.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
Fig. 13.33 Sketch to Problem 13.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
Fig. 13.34 Sketch to Problem 13.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Fig. 13.35 Sketch to Problem 13.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
Fig. 13.36 Solution to Problem 13.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Fig. 13.37 Sketch to Problem 13.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
Fig. 13.38 Sketch to Problem 13.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Fig. 13.39 Sketch to Problem 13.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
Fig. 13.40 Sketch to Problem 13.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
Fig. 13.41 Sandcastle according to [2] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
Fig. 14.1 Wire pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336
Fig. 14.2 Sketch to Problem 14.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
Fig. 14.3 Wall—Heat conduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
Fig. 14.4 Heat transfer between two homogeneous systems . . . . . . . . . . . 344
Fig. 14.5 Thermal balancing process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
Fig. 14.6 Sketch to Problem 14.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
xxviii List of Figures

Fig. 14.7 Sketch to Problem 14.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350


Fig. 14.8 Sketch to Problem 14.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
Fig. 15.1 Contradiction to the second law according to Planck . . . . . . . . . 356
Fig. 15.2 Thermal engine (closed system) without cooling . . . . . . . . . . . . 357
Fig. 15.3 Thermal engine (open system inside) in black box notation . . . 358
Fig. 15.4 Thermal engine (closed system) in a p, V -diagram . . . . . . . . . . 361
Fig. 15.5 Thermal engine (open system) in a p, v-diagram . . . . . . . . . . . . 362
Fig. 15.6 Contradiction to the second law according to Clausius . . . . . . . 363
Fig. 15.7 Cooling machine/heat pump in black box notation . . . . . . . . . . 365
Fig. 15.8 Temperature gradient clarification . . . . . . . . . . . . . . . . . . . . . . . . 366
Fig. 15.9 Cooling machine/heat pump (closed system)
in a p, V -diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
Fig. 15.10 Solution to Problem 15.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
Fig. 15.11 Clockwise Carnot process—thermal engine . . . . . . . . . . . . . . . . 373
Fig. 15.12 Clockwise Carnot process— T, s-diagram . . . . . . . . . . . . . . . . . 374
Fig. 15.13 Counterclockwise Carnot process—cooling machine/heat
pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376
Fig. 16.1 Evaluation of thermal energy sources . . . . . . . . . . . . . . . . . . . . . 380
Fig. 16.2 Gaining work from heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
Fig. 16.3 Energy-in-exergy-and-anergy-decomposition-machine . . . . . . . 382
Fig. 16.4 Heat at constant temperature → Evaluation of the exergy
of the heat q@T with a Carnot-machine . . . . . . . . . . . . . . . . . . . 383
Fig. 16.5 Exergy of heat at isothermal expansion according to [3] . . . . . . 384
Fig. 16.6 Non-isothermal heat supply (1)→(2) . . . . . . . . . . . . . . . . . . . . . 385
Fig. 16.7 Exergy of heat—influence of the environmental
temperature according to [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
Fig. 16.8 Sign of exergy of heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
Fig. 16.9 Exergy of a fluid flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
Fig. 16.10 Change of state (1)→(env) with p1 > penv , according
to [3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
Fig. 16.11 Change of state (1)→(env) with p1 < penv , according
to [3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
Fig. 16.12 Exergy of a closed system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
Fig. 16.13 Change of state (1)→(env) with p1 > penv . . . . . . . . . . . . . . . . 394
Fig. 16.14 Change of state (1)→(env) with p1 < penv . . . . . . . . . . . . . . . . 396
Fig. 16.15 Closed system: balance of exergy . . . . . . . . . . . . . . . . . . . . . . . . 399
Fig. 16.16 Overview closed system: balance of energy (a), entropy
(b) and exergy (c) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400
Fig. 16.17 Open system in steady state: balance of exergy . . . . . . . . . . . . . 402
Fig. 16.18 Overview open system in steady state: balance of energy
(a), entropy (b) and exergy (c) . . . . . . . . . . . . . . . . . . . . . . . . . . . 404
Fig. 16.19 Balance of exergy: a clockwise cycle, b counterclockwise
cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
Fig. 16.20 Overview clockwise cycle: balance of energy (a), entropy
(b) and exergy (c) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
List of Figures xxix

Fig. 16.21 Overview counterclockwise cycle: balance of energy


(a), entropy (b) and exergy (c) . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
Fig. 16.22 Illustration of energy fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
Fig. 16.23 Sankey-diagram of an open system . . . . . . . . . . . . . . . . . . . . . . . 413
Fig. 16.24 Heat transfer (wall) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414
Fig. 16.25 Sketch for the solution to Problem 16.2 . . . . . . . . . . . . . . . . . . . 415
Fig. 16.26 Sketch for the solution to Problem 16.2 . . . . . . . . . . . . . . . . . . . 417
Fig. 16.27 Sketch for the solution to Problem 16.3 . . . . . . . . . . . . . . . . . . . 418
Fig. 16.28 Sketch for the solution to Problem 16.3 . . . . . . . . . . . . . . . . . . . 421
Fig. 17.1 Turbine: a First law, b Second law, c Exergy . . . . . . . . . . . . . . . 424
Fig. 17.2 Adiabatic turbine: Illustration in a T, s-diagram (ideal
and real gases) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426
Fig. 17.3 Compressor: a First law, b Second law, c Exergy . . . . . . . . . . . . 428
Fig. 17.4 Adiabatic compressor: Illustration in a T, s-diagram
(ideal and real gases) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
Fig. 17.5 Thermal turbo-machines in a h, s-diagram . . . . . . . . . . . . . . . . . 431
Fig. 17.6 Adiabatic throttle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433
Fig. 17.7 Symbols for heat exchangers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
Fig. 17.8 Counterflow heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
Fig. 17.9 Counterflow heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437
Fig. 17.10 Temperature profiles in a heat exchanger: a Parallel flow
b Counterflow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
Fig. 17.11 Entropy balance (heat exchanger is supposed to be
adiabatic to the environment) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
Fig. 17.12 Entropy balance (system C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
Fig. 17.13 Overall entropy balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441
Fig. 17.14 Carnot process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
Fig. 17.15 Joule process—layout clockwise cycle . . . . . . . . . . . . . . . . . . . . 444
Fig. 17.16 Joule process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
Fig. 17.17 Clausius Rankine process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
Fig. 17.18 Clausius Rankine process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 446
Fig. 17.19 Seiliger process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
Fig. 17.20 Stirling process—possible technical principle . . . . . . . . . . . . . . 449
Fig. 17.21 Stirling process—p, v- and T, s-diagram . . . . . . . . . . . . . . . . . . . 449
Fig. 17.22 Compression heat pump—layout . . . . . . . . . . . . . . . . . . . . . . . . . 453
Fig. 17.23 Compression heat pump: sketch and illustration
in a T, s-diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
Fig. 17.24 Sketch of the layout to Problem 17.1 . . . . . . . . . . . . . . . . . . . . . . 456
Fig. 17.25 p, v- and T, s-diagram to Problem 17.1 . . . . . . . . . . . . . . . . . . . . 456
Fig. 17.26 Sketch to Problem 17.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
Fig. 17.27 Sketch to Problem 17.2: a Compressor in a p, v-diagram,
b Heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Fig. 17.28 Sketch to Problem 17.2(c) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Fig. 17.29 Sketch to Problem 17.2(e) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463
Fig. 17.30 Sketch to Problem 17.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
xxx List of Figures

Fig. 17.31 Sketch to Problem 17.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468


Fig. 18.1 Compressibility factor Z for air according to [4] . . . . . . . . . . . . 475
Fig. 18.2 Changes of aggregation state . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
Fig. 18.3 Isobaric vaporisation—schematic illustration . . . . . . . . . . . . . . . 477
Fig. 18.4 Isobaric vaporisation—schematic illustration in a p,
v-diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479
Fig. 18.5 The p, v, T-diagram: a Real fluid, b Ideal gas . . . . . . . . . . . . . . . 481
Fig. 18.6 The p, v, T-diagram: examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
Fig. 18.7 Possible projections of the p, v, T-state space . . . . . . . . . . . . . . . 484
Fig. 18.8 The p, T-diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
Fig. 18.9 The p, T-diagram: a Water, b Non-water fluid . . . . . . . . . . . . . . 486
Fig. 18.10 The T, v-diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487
Fig. 18.11 The p, v-diagram for water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 488
Fig. 18.12 Closed system—Isothermal expansion . . . . . . . . . . . . . . . . . . . . 488
Fig. 18.13 Isothermal expansion (1) to (4) in a p, v-diagram . . . . . . . . . . . . 489
Fig. 18.14 Lever rule of the quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 492
Fig. 18.15 Van der Waals approach for CO2 . . . . . . . . . . . . . . . . . . . . . . . . . 496
Fig. 18.16 Isobaric vaporisation: a reversible, b irreversible . . . . . . . . . . . . 505
Fig. 18.17 Isobaric vaporisation of water . . . . . . . . . . . . . . . . . . . . . . . . . . . 508
Fig. 18.18 Specific enthalpy of vaporisation h v of water . . . . . . . . . . . . . 509
Fig. 18.19 T, s-diagram of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
Fig. 18.20 T, s-diagram of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
Fig. 18.21 h, s-diagram of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
Fig. 18.22 log p, h-diagram of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
Fig. 18.23 Technical work in the h, s-diagram . . . . . . . . . . . . . . . . . . . . . . . 515
Fig. 18.24 Isobaric, isothermal, reversible change of state . . . . . . . . . . . . . . 516
Fig. 18.25 Isochoric change of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
Fig. 18.26 Adiabatic change of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
Fig. 18.27 Clausius-Clapeyron relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519
Fig. 18.28 Vapour pressure curve of water . . . . . . . . . . . . . . . . . . . . . . . . . . 520
Fig. 18.29 Adiabatic throttling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522
Fig. 18.30 Adiabatic throttling (Joule-Thomson effect) . . . . . . . . . . . . . . . . 525
Fig. 18.31 Joule-Thomson coefficient δh for several gases . . . . . . . . . . . . . 526
Fig. 18.32 Joule-Thomson coefficient δh for air . . . . . . . . . . . . . . . . . . . . . . 526
Fig. 18.33 Adiabatic throttling of air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
Fig. 18.34 Solution to Problem 18.3(e) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
Fig. 18.35 Solution to Problem 18.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533
Fig. 19.1 Mixture of gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 538
Fig. 19.2 Adiabatic mixing of two gases according to Dalton . . . . . . . . . . 541
Fig. 19.3 Gas constant of a mixture of ideal gases . . . . . . . . . . . . . . . . . . . 547
Fig. 19.4 Adiabatic mixing closed system—exemplary for n = 3
components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 550
Fig. 19.5 Adiabatic mixing open system—exemplary for n = 3
components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
Fig. 19.6 Irreversibility of mixing (irreversible, adiabatic) . . . . . . . . . . . . 553
List of Figures xxxi

Fig. 19.7 Irreversibility of mixing—equivalent model (reversible,


diabatic) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553
Fig. 19.8 Thought experiment for a reversible, isothermal mixing:
pistons replace separating wall in Fig. 19.2 . . . . . . . . . . . . . . . . . 553
Fig. 19.9 Sketch to Problem 19.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
Fig. 20.1 Humid air—a unsaturated, b saturated, c saturated
with additional liquid/ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
Fig. 20.2 p, T -diagram of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
Fig. 20.3 Distinction between a vaporisation and b evaporation . . . . . . . . 569
Fig. 20.4 Humid air—Example 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
Fig. 20.5 Humid air—Example 2/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575
Fig. 20.6 How to treat the water in humid air—vapour . . . . . . . . . . . . . . . 584
Fig. 20.7 How to treat the water in humid air—liquid . . . . . . . . . . . . . . . . 585
Fig. 20.8 How to treat the water in humid air—ice . . . . . . . . . . . . . . . . . . 587
Fig. 20.9 How to treat the water in humid air—vapour . . . . . . . . . . . . . . . 589
Fig. 20.10 How to treat the water in humid air—liquid . . . . . . . . . . . . . . . . 590
Fig. 20.11 How to treat the water in humid air—ice . . . . . . . . . . . . . . . . . . 591
Fig. 20.12 Unsaturated humid air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 592
Fig. 20.13 Saturated humid air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 593
Fig. 20.14 Saturated humid air and liquid water . . . . . . . . . . . . . . . . . . . . . . 594
Fig. 20.15 Saturated humid air and solid water . . . . . . . . . . . . . . . . . . . . . . . 594
Fig. 20.16 Saturated humid air, liquid and solid water . . . . . . . . . . . . . . . . . 595
Fig. 20.17 Sketch to Problem 20.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 596
Fig. 20.18 Solution to Problem 20.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 597
Fig. 20.19 Solution to Problem 20.2—Phase change of the water . . . . . . . . 599
Fig. 20.20 Mollier h 1+x , x-diagram—schematic . . . . . . . . . . . . . . . . . . . . . 603
Fig. 20.21 Heating (1) → (2) and cooling (1) → (3) at constant
water content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605
Fig. 20.22 Dehumidification of air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
Fig. 20.23 Adiabatic mixing of humid air . . . . . . . . . . . . . . . . . . . . . . . . . . . 610
Fig. 20.24 Adiabatic mixing of humid air, h 1+x , x-diagram . . . . . . . . . . . . 611
Fig. 20.25 Adiabatic mixing of humid air, h 1+x , x-diagram . . . . . . . . . . . . 613
Fig. 20.26 Humidification of air with pure water . . . . . . . . . . . . . . . . . . . . . 614
Fig. 20.27 Humidification of air with pure water . . . . . . . . . . . . . . . . . . . . . 616
Fig. 20.28 Adiabatic saturation temperature—Thermodynamic
equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 618
Fig. 20.29 Adiabatic saturation temperature . . . . . . . . . . . . . . . . . . . . . . . . . 621
Fig. 20.30 Psychrometer (wet-and-dry-bulb thermometer) . . . . . . . . . . . . . 622
Fig. 20.31 The h 1+x , x-diagram— p = 1 bar . . . . . . . . . . . . . . . . . . . . . . . . 623
Fig. 20.32 Solution to Problem 20.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
Fig. 20.33 Technical layout to Problem 20.4 . . . . . . . . . . . . . . . . . . . . . . . . . 626
Fig. 20.34 Solution to Problem 20.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
Fig. 21.1 Steady state flow process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
Fig. 21.2 Incompressible flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
Fig. 21.3 Adiabatic flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635
xxxii List of Figures

Fig. 21.4 Adiabatic flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637


Fig. 21.5 Adiabatic diffusor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
Fig. 21.6 Adiabatic diffusor in a h, s-diagram . . . . . . . . . . . . . . . . . . . . . . 640
Fig. 21.7 Adiabatic nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
Fig. 21.8 Adiabatic nozzle in a h, s-diagram . . . . . . . . . . . . . . . . . . . . . . . 640
Fig. 21.9 Velocity of sound a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
Fig. 21.10 Adiabatic, frictional tube flow ( A = const.) . . . . . . . . . . . . . . . . 648
Fig. 21.11 Subsonic Fanno-curves, ideal gas with κ = const. . . . . . . . . . . . 651
Fig. 21.12 Subsonic Fanno-curves for one mass flux density, ideal
gas with κ = const. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 651
Fig. 21.13 Supersonic Fanno-curves, ideal gas with κ = const. . . . . . . . . . 654
Fig. 21.14 Supersonic Fanno-curves for one mass flux density, ideal
gas with κ = const. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655
Fig. 21.15 Non-adiabatic, frictionless tube flow ( A = const.) . . . . . . . . . . . 659
Fig. 21.16 Subsonic Rayleigh-curves for heating, ideal gas
with κ = const. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
Fig. 21.17 Subsonic Rayleigh-curves for one mass flux density
and heating, ideal gas with κ = const. . . . . . . . . . . . . . . . . . . . . . 662
Fig. 21.18 Rayleigh correlation (subsonic), p, v-diagram . . . . . . . . . . . . . . 663
Fig. 21.19 Polytropic exponent—a subsonic, b supersonic . . . . . . . . . . . . . 664
Fig. 21.20 Supersonic Rayleigh-curves for heating, ideal gas
with κ = const. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
Fig. 21.21 Supersonic Rayleigh-curves for one mass flux density
and heating, ideal gas with κ = const. . . . . . . . . . . . . . . . . . . . . . 665
Fig. 21.22 Rayleigh correlation (supersonic), p, v-diagram . . . . . . . . . . . . . 666
Fig. 21.23 Normal shock, adiabatic tube flow . . . . . . . . . . . . . . . . . . . . . . . . 669
Fig. 21.24 Steady flow from vessels according to [1] . . . . . . . . . . . . . . . . . . 670
Fig. 21.25 Local flow function  for κ = 1.4, state (0) in rest . . . . . . . . . . 673
Fig. 21.26 Converging nozzle scenarios I/II . . . . . . . . . . . . . . . . . . . . . . . . . 675
Fig. 21.27 Converging nozzle scenarios II/II . . . . . . . . . . . . . . . . . . . . . . . . 676
Fig. 21.28 Overview subsonic/supersonic flows . . . . . . . . . . . . . . . . . . . . . . 681
Fig. 21.29 Steady flow from vessels using a Laval-nozzle . . . . . . . . . . . . . . 682
Fig. 21.30 Pressure characteristic Laval-nozzle . . . . . . . . . . . . . . . . . . . . . . 682
Fig. 21.31 a Venturi-nozzle, f Supersonic flow . . . . . . . . . . . . . . . . . . . . . . 683
Fig. 21.32 Supersonic exit Laval-nozzle a isentropic,
b polytropic/adiabatic, i.e. with friction . . . . . . . . . . . . . . . . . . . 686
Fig. 21.33 Supersonic exit Laval-nozzle, T, s-diagram . . . . . . . . . . . . . . . . 687
Fig. 21.34 Laval-diffusor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
Fig. 22.1 Clockwise Carnot process: thermal engine . . . . . . . . . . . . . . . . . 690
Fig. 22.2 Clausius–Rankine cycle: sketch . . . . . . . . . . . . . . . . . . . . . . . . . . 691
Fig. 22.3 Clausius–Rankine cycle: a p, v-diagram, b T, s-diagram . . . . . 691
Fig. 22.4 Steam power process: boiler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694
Fig. 22.5 Steam power process: T, s-diagram . . . . . . . . . . . . . . . . . . . . . . . 694
Fig. 22.6 Steam power process: h, s-diagram . . . . . . . . . . . . . . . . . . . . . . . 695
Fig. 22.7 Steam power process—expansion in the wet-steam region . . . . 697
List of Figures xxxiii

Fig. 22.8 Steam power process—reheating . . . . . . . . . . . . . . . . . . . . . . . . . 697


Fig. 22.9 Steam power process—reheating . . . . . . . . . . . . . . . . . . . . . . . . . 698
Fig. 22.10 Steam power process—a No feed water preheating
b Regenerative feed water preheating, cf. Fig. 22.11 . . . . . . . . . 700
Fig. 22.11 Steam power process—regenerative feed water preheating . . . . 700
Fig. 22.12 Compression heat pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
Fig. 22.13 Compression heat pump—layout (left), T, s-diagram
(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
Fig. 22.14 Compression heat pump—log p, h-diagram . . . . . . . . . . . . . . . 705
Fig. 22.15 Absorption heat pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 706
Fig. 22.16 Solar cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707
Fig. 22.17 Sketch to Problem 22.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707
Fig. 22.18 h, s-diagram to Problem 22.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 708
Fig. 22.19 Entropy balance to Problem 22.1 . . . . . . . . . . . . . . . . . . . . . . . . . 710
Fig. 22.20 a Sketch of the plant, b T, s-diagram to Problem 22.2 . . . . . . . . 713
Fig. 22.21 a Energy balance, b Entropy balance for the condenser . . . . . . . 715
Fig. 23.1 Fossil fuels according to [6] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 720
Fig. 23.2 Combustion process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 722
Fig. 23.3 Example of a solid fossil fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725
Fig. 23.4 Exemplary gaseous fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 732
Fig. 23.5 Exemplary gaseous fuel mixture . . . . . . . . . . . . . . . . . . . . . . . . . 743
Fig. 23.6 Condensing product water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 747
Fig. 23.7 Energy balance—combustion chamber . . . . . . . . . . . . . . . . . . . . 757
Fig. 23.8 Definition of the lower heating value HU . . . . . . . . . . . . . . . . . . 759
Fig. 23.9 Combustion—Step 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 761
Fig. 23.10 Combustion—Step 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 762
Fig. 23.11 Combustion—Step 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 762
Fig. 23.12 Combustion—summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 764
Fig. 23.13 Condensation of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 764
Fig. 23.14 HU (T0 ) → HU (T ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 767
Fig. 23.15 Combustion with unsaturated humid air—summary . . . . . . . . . 768
Fig. 23.16 Combustion with saturated humid air—summary . . . . . . . . . . . 769
Fig. 23.17 Definition of the higher heating value H0 . . . . . . . . . . . . . . . . . . 771
Fig. 23.18 Combustion with saturated humid air—applying
the specific higher heating value . . . . . . . . . . . . . . . . . . . . . . . . . 774
Fig. 23.19 Component-wise calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 776
Fig. 23.20 h, ϑ-diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 781
Fig. 23.21 Adiabatic flame temperature (dry air) . . . . . . . . . . . . . . . . . . . . . 782
Fig. 23.22 Combustion chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784
Fig. 23.23 Combustion chamber—variation of λ . . . . . . . . . . . . . . . . . . . . . 786
Fig. 23.24 Influence of fuel-air equivalence ratio and air preheating
on the adiabatic flame temperature for oil according to [7] . . . . 786
Fig. 23.25 Solution to Problem 23.5b . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 788
Fig. 23.26 Sketch to Problem 23.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 792
Fig. 24.1 Enthalpy of reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 810
xxxiv List of Figures

Fig. 24.2 Enthalpy of reaction at T0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 811


Fig. 24.3 Theoretical adiabatic flame temperature of methane . . . . . . . . . 814
Fig. 24.4 Entropy and irreversibility of a reaction . . . . . . . . . . . . . . . . . . . 814
Fig. 24.5 First law of thermodynamics reactive closed system . . . . . . . . . 815
Fig. 24.6 Second law of thermodynamics reactive closed system . . . . . . . 816
Fig. 24.7 Sketch to Problem 24.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 817
Fig. 24.8 Gibbs energy—closed system . . . . . . . . . . . . . . . . . . . . . . . . . . . 822
Fig. 24.9 Gibbs energy—open system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824
Fig. 24.10 Mixing of two ideal gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 830
Fig. 24.11 Chemical potential at evaporation . . . . . . . . . . . . . . . . . . . . . . . . 833
Fig. 24.12 Dynamic chemical reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 835
Fig. 24.13 Chemical balancing driven by μi . . . . . . . . . . . . . . . . . . . . . . . . . 836
Fig. 24.14 Exergy of a fossil fuel—energy balance . . . . . . . . . . . . . . . . . . . 837
Fig. 24.15 Exergy of a fossil fuel—entropy balance . . . . . . . . . . . . . . . . . . 837
Fig. 24.16 Exergy of a fossil fuel—exergy balance . . . . . . . . . . . . . . . . . . . 839
Fig. C.1 log p, h-diagram of water, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 950
Fig. C.2 T, s-diagram of water, generated with XSteam, see [9] . . . . . . . 951
Fig. C.3 h, s-diagram of water, generated with XSteam, see [9] . . . . . . . 952
Fig. C.4 log p, h-diagram of R134a, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 954
Fig. C.5 log p, h-diagram of R290, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 955
Fig. C.6 log p, h-diagram of R717, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 956
Fig. C.7 log p, h-diagram of R744, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 957
Fig. C.8 log p, h-diagram of R1234yf, generated with CoolProp,
see [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 958
Fig. D.1 h 1+x , x-diagram (atmospheric air + water) . . . . . . . . . . . . . . . . 960
Fig. D.2 h 1+x , x-diagram (hydrogen + water) . . . . . . . . . . . . . . . . . . . . . . 961
List of Tables

Table 3.1 Temperature measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47


Table 3.2 Molar masses of relevant elements . . . . . . . . . . . . . . . . . . . . . . . 54
Table 5.1 Exemplary combinations of state values for single ideal
gases in thermodynamic equilibrium . . . . . . . . . . . . . . . . . . . . . 73
Table 12.1 Isentropic exponent for ideal gases at standard
conditions, i.e. temperature ϑ = 0 ◦ C according to DIN
1343, see [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Table 12.2 Temperature dependence of an ideal gas according
to Problem 12.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
Table 13.1 Polytropic change of state—Technical meaning for ideal
gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
Table 17.1 Overview thermodynamic cycles . . . . . . . . . . . . . . . . . . . . . . . . 455
Table 18.1 Triple/critical point of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486
Table 18.2 Further equations of state for real fluids . . . . . . . . . . . . . . . . . . . 493
Table 20.1 Properties—Humid air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 592
Table 23.1 Simplified composition of dry, atmospheric air . . . . . . . . . . . . . 727
Table 23.2 Lower heating value HU (solids and liquids)
at ϑ0 = 25 ◦ C, see [10] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 760
Table 23.3 Lower heating value HU (Gases) at ϑ0 = 25 ◦ C, see [10] . . . . . 760
Table 23.4 Averaged specific heat capacity c p |ϑ0 in kgkJK , according
to [5] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 766
Table 23.5 Higher heating values H0 at ϑ0 = 25 ◦ C, according to [10] . . . 773
Table 23.6 Higher heating values H0 (Gases) at ϑ0 = 25 ◦ C,
according to [10] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774
Table 23.7 Composition of the combustion air . . . . . . . . . . . . . . . . . . . . . . . 789
Table 23.8 Averaged specific heat capacities c p |ϑ0 in kgkJK . . . . . . . . . . . . . 789
Table 23.9 Iterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 792
Table 24.1 Chemical composition of a standard atmosphere at 25 ◦ C
and 1 bar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 837
Table A.1 Saturated liquid and saturated steam 1/4 . . . . . . . . . . . . . . . . . . 846

xxxv
xxxvi List of Tables

Table A.2 Saturated liquid and saturated steam 2/4 . . . . . . . . . . . . . . . . . . 847


Table A.3 Saturated liquid and saturated steam 3/4 . . . . . . . . . . . . . . . . . . 849
Table A.4 Saturated liquid and saturated steam 4/4 . . . . . . . . . . . . . . . . . . 852
3
Table A.5 Specific volume v of water in mkg 1/2 . . . . . . . . . . . . . . . . . . . . . 854
3
Table A.6 Specific volume v of water in mkg 2/2 . . . . . . . . . . . . . . . . . . . . . 856
kJ
Table A.7 Specific enthalpy h of water in kg 1/2 . . . . . . . . . . . . . . . . . . . . 858
kJ
Table A.8 Specific enthalpy h of water in kg 2/2 . . . . . . . . . . . . . . . . . . . . 860
Table A.9 Specific entropy s of water in kgkJK 1/2 . . . . . . . . . . . . . . . . . . . . 862
kJ
Table A.10 Specific entropy s of water in kgK 2/2 . . . . . . . . . . . . . . . . . . . . 864
Table A.11 Specific heat capacity c p of water in kgkJK 1/2 . . . . . . . . . . . . . . 866
Table A.12 Specific heat capacity c p of water in kgkJK 2/2 . . . . . . . . . . . . . . 868
Table B.1 Absolute molar specific enthalpy and entropy of H2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874
Table B.2 Absolute molar specific enthalpy and entropy of H
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 877
Table B.3 Absolute molar specific enthalpy and entropy of O2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
Table B.4 Absolute molar specific enthalpy and entropy of O
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 883
Table B.5 Absolute molar specific enthalpy and entropy of OH
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886
Table B.6 Absolute molar specific enthalpy and entropy
of H2 O(liq). The dependency of the enthalpy
on the pressure is supposed to be insignificant.
Liquid water is treated as an incompressible fluid, i.e.
the entropy does not need to be corrected. Reference
temperature for averaged heat capacities is ϑ0 = 0 ◦ C . . . . . . . 889
Table B.7 Absolute molar specific enthalpy and entropy of H2 O(g)
at p0 = 1 bar. Vapour is treated as an ideal gas.
Reference temperature for averaged heat capacities is
ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 890
Table B.8 Absolute molar specific enthalpy and entropy of N2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 893
Table B.9 Absolute molar specific enthalpy and entropy of N
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 896
List of Tables xxxvii

Table B.10 Absolute molar specific enthalpy and entropy of NO


at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 899
Table B.11 Absolute molar specific enthalpy and entropy of NO2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 902
Table B.12 Absolute molar specific enthalpy and entropy of CO
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905
Table B.13 Absolute molar specific enthalpy and entropy of CO2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 908
Table B.14 Absolute molar specific enthalpy and entropy of CGraphite
The dependency of the enthalpy on the pressure
is supposed to be insignificant. Graphite is treated
as an incompressible solid, i.e. the entropy does not need
to be corrected. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 911
Table B.15 Absolute molar specific enthalpy and entropy of S
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914
Table B.16 Absolute molar specific enthalpy and entropy of S(a)
The dependency of the enthalpy on the pressure is
supposed to be insignificant. Sulphur(a) is treated
as an incompressible solid, i.e. the entropy does not need
to be corrected. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
Table B.17 Absolute molar specific enthalpy and entropy of S(b)
The dependency of the enthalpy on the pressure is
supposed to be insignificant. Sulphur(b) is treated
as an incompressible solid, i.e. the entropy does not need
to be corrected . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 918
Table B.18 Absolute molar specific enthalpy and entropy of S(liq)
The dependency of the enthalpy on the pressure is
supposed to be insignificant. Sulphur(liq) is treated
as an incompressible liquid, i.e. the entropy does
not need to be corrected . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 919
Table B.19 Absolute molar specific enthalpy and entropy of S2
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 920
Table B.20 Absolute molar specific enthalpy and entropy of SO
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 923
xxxviii List of Tables

Table B.21 Absolute molar specific enthalpy and entropy of SO2


at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 926
Table B.22 Absolute molar specific enthalpy and entropy of CH4
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 929
Table B.23 Absolute molar specific enthalpy and entropy of C2 H6
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 932
Table B.24 Absolute molar specific enthalpy and entropy of C3 H8
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935
Table B.25 Absolute molar specific enthalpy and entropy
of C2 H5 OH at p0 = 1 bar. Reference temperature
for averaged heat capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . 938
Table B.26 Absolute molar specific enthalpy and entropy
of C2 H5 OH(liq). The dependency of the enthalpy
on the pressure is supposed to be insignificant. Liquid
ethanol is treated as an incompressible fluid, i.e.
the entropy does not need to be corrected. Reference
temperature for averaged heat capacities is ϑ0 = 0 ◦ C . . . . . . . 941
Table B.27 Absolute molar specific enthalpy and entropy of CH3 OH
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 942
Table B.28 Absolute molar specific enthalpy and entropy
of CH3 OH(liq). The dependency of the enthalpy
on the pressure is supposed to be insignificant. Liquid
methanol is treated as an incompressible fluid, i.e.
the entropy does not need to be corrected. Reference
temperature for averaged heat capacities is ϑ0 = 0 ◦ C . . . . . . . 945
Table B.29 Absolute molar specific enthalpy and entropy of air
at p0 = 1 bar. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦ C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946
Chapter 1
Introduction

Thermodynamics is one of the most difficult and probably even one of the most
challenging disciplines in mechanical engineering. However, thermodynamics is
necessary to understand the principles of energy conversion, e.g. when designing
thermal machines such as gas turbines, internal combustion engines or even fuel cells
in modern technical applications. The fundamental principles of energy conversion
are actually what thermodynamics is about: It is well known from physics lessons,
that energy is conserved, i.e. energy can not be generated or destroyed. This is what
the first law of thermodynamics summarises. Furthermore, energy conversion, e.g.
from thermal energy to mechanical energy, is limited. A conventional combustion
engine converts chemical bonded energy into mechanical energy, but as we know,
part of this converted energy has to be released to the environment, e.g. by a cooler.
This limitation, however, obeys the second law of thermodynamics.
Although the fundamentals of thermodynamics date back to the 18th century,
their principles still apply today. The word thermodynamics is composed of the
Greek words therme (heat) and dynamis (force): It covers the theory of of energy
and its convertibility. Even though the word dynamics suggests that systems are in
motion, this book deals with so-called equilibrium states.
As already mentioned, the conversion of energy plays an important role for
mechanical engineers today. Examples from everyday life are:
• Vehicles with internal combustion engines convert chemical bonded energy of the
fuel to mechanical energy for the powertrain. Combustion processes still represent
mobility at the beginning of the 21st century. For various reasons, it makes sense
to strengthen renewable energies and make electric driving possible. New tech-
nologies, in particular electric storage systems, are needed to supply the electric
drives with energy. It is also possible to recover energy, e.g. when braking the vehi-
cle: Instead of transferring mechanical energy into heat at the braking disks, this
mechanical energy is converted into electrical energy for recharging a battery. The
conversion of energy thus also plays a decisive role in new modern drive systems.

© Springer Nature Switzerland AG 2022 1


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_1
2 1 Introduction

• Thermal power plants fire fossil fuels in order to generate electricity. Hence, chem-
ically bound energy, i.e. fossil fuels, is converted into thermal energy in the com-
bustion chamber of a power plant. This thermal energy can then be converted into
electrical energy in the turbine generator unit.
• Solar thermal power plants utilise solar energy to generate heat, which is converted
into electricity via turbines.

All these examples are intended to show, that thermodynamics, as an important


basic subject, is indispensable for the energy turnaround that has already been ini-
tiated. The sustainable energy supply of a constantly growing global population is
one of the future megatrends. The aim of thermodynamics is to provide the laws of
energy conversion. In addition, engineers are to be enabled to develop and design
technical applications for future energy conversion machines.

1.1 How Is This Book Structured?

The structure of the book is derived from the module descriptions that can be found
at German universities in the field of thermodynamics. Due to the complexity of the
subject, the lecture is divided into two parts, namely Technical Thermodynamics I
and Technical Thermodynamics II. However, I have decided to divide the book into
three parts.
Part I deals with the basics of the topic: A conceptual clarification of thermody-
namic systems and their states is given. Simple systems are investigated in thermo-
dynamic equilibrium. Heat and work as process variables are introduced, which can
drive a system into a new state of equilibrium. The first law of thermodynamics, the
energy conservation principle that every system must follow, is applied. This princi-
ple makes it possible to understand energy transformations, e.g. from thermal energy
to mechanical energy in a so-called thermal engine. In order to optimise the conver-
sion rate, the second law of thermodynamics and thus entropy as a new state variable
is indispensable. Thermodynamic cycles are included in Part I as well. Liquids in
Part I are ideal, i.e. they behave either like an ideal gas or like an incompressible
liquid. The thermodynamics of transient processes, e.g. the cooling of a system to a
new state of equilibrium, is also covered with in Part I. The didactics of the first part
follow Erich Hahne, see [1], whose textbook is excellently suited and designed for
access to the subject of technical thermodynamics. I tried to modify the structure of
the first chapters until I realised that his structure was already perfect—so I decided
to leave it unchanged. It was also Erich Hahne who inspired me through personal
conversations during my first semester at Hamburg University of Applied Sciences.
I am therefore very grateful to him.
While liquids in Part I are considered as ideal gases or incompressible liquids,
Part II deals with fluids that show deviations from an idealised behaviour and that
additionally can change their state of aggregation. For example, water can change
its state of aggregation by supplying or releasing thermal energy, e.g. it freezes by
1.1 How Is This Book Structured? 3

cooling so that liquid water becomes a solid. The physical properties, i.e. the state
values, vary with each change in the aggregate state. These so-called real fluids do no
longer obey the thermal/caloric equations of state, that have been applied in Part I for
ideal gases. The task of thermodynamics is to quantify the state of the system and to
provide means for calculating the change of state. Part II also includes gas mixtures,
i.e. the presence of several gaseous components without being chemically reactive.
An example of a gas mixture is humid air as a mixture of dry air and vaporous water.
The amount of steam in humid air depends on its temperature: as the temperature
drops, liquid water or even ice can form. Clockwise or counterclockwise thermo-
dynamic cycles, which are operated with real fluids, are introduced and calculated.
Supersonic flows, usually part of the field of Gas Dynamics, are briefly discussed.
Part III finally implies chemically reactive systems and will complete this text-
book. Focus will be on fossil fuels: When fossil fuels, i.e. fuels containing hydrogen
and carbon, are combusted with air, carbon dioxide and water are produced,1 i.e. not
only an energetic conversion takes place but additionally a conversion of substances.
Thus, the resulting exhaust gas is a mixture of various components, as discussed in
Part II. For the design of a combustion process, it is important to determine the state
of aggregation of the water in the exhaust gas. Condensed water for instance can
easily lead to corrosive liquids, which cause several problems in technical systems.
Furthermore, the specific lower/higher heating value is utilised to predict exhaust
gas temperature as well as released heat. An alternative approach, following the
principles of absolute enthalpy/entropy, is investigated in the last section of Part III.
This approach allows to describe any chemical reaction energetically and to apply
the second law of thermodynamics in order to specify the irreversibility of chemical
reactions. The application of this approach furthermore enables the investigation of
modern energy conversion systems, such as fuel cell systems.
Finally, Fig. 1.1 shows the classical fields of conventional technical thermody-
namics. All three fields are part of this book and are introduced briefly below.
This books contains several examples and exercises. However, from time to time,
especially in Part I, some of these examples require the knowledge of upcoming
chapters. Don’t let that demotivate you. Access to the questions is obtained by a
second study of the book after the basic philosophy of thermodynamics is understood.

State of a system/Phase change

It turns out that it is imperative to define a system and its state before energy con-
version can be studied. A system can be open or closed with regard to the exchange
of mass across system boundaries. Furthermore, energy in form of work or heat can
cross the system boundary. In any case, the state of a system can be modified by
external influences.

1The exhaust gas may also consist of nitrogen, in case the air contains nitrogen, and oxygen, in
case more air is supplied than required for a stoichiometric combustion.
4 1 Introduction

Fig. 1.1 What is thermodynamics all about? (Tasks of thermodynamics)

As an example, the state of a mass in a gravitational field can be described by


its temperature and its position in the gravitational field. By lifting the mass, i.e. by
an external intervention, the vertical coordinate changes. As a result, the potential
energy has increased. If this mass falls to the ground, it is subjected to an acceleration
and potential energy is converted into kinetic energy. When the mass finally hits the
ground, its temperature rises.
For more complex systems, not only the temperature or the vertical position are
crucial for the state description, but also other additional information such as the
pressure, which is necessary for an accurate description of the system. Besides solids,
a system can also consist of gaseous fluids. It is known from everyday experience
that gases can be compressed as the pressure increases, see Fig. 1.1. The state values
of pressure and temperature are functions of the mass and volume of a gas and, in
an idealised case, obey the gas equation. Furthermore, at low temperatures and high
pressures, gases can also be liquefied. Unfortunately, such a phase change cannot be
treated with the theory of ideal gases any more and is discussed in Part II.

Changeability of energy

The physical principle of conservation of energy, considered as the first law of ther-
modynamics, is introduced and applied to calculate the conversion of energy from
one form to another. For example, an idealised pendulum converts potential energy
into kinetic energy and vice versa. Figure 1.1 also shows another possible conversion
1.1 How Is This Book Structured? 5

of energy. Heat, i.e. thermal energy, is supplied to the system so that the gas in the
cylinder expands and the spring mounted on the outside is compressed. Energy is
therefore now present as mechanical energy in the spring.
In Part I, the limitations to which energy conversion is subject are derived. It
is shown that it is impossible to design a machine that, for example, completely
converts heat into work: It is common knowledge that a coal-fired power plant cannot
operate without cooling towers, which obviously release thermal energy into the
environment. Such cooling reduces the benefit of such a power plant, as the chemical
energy of the fuel is not fully converted into electricity.
For an engineer, the question inevitably arises as to the maximum efficiency at
which an ideal power plant can be operated. The derivation of this maximum effi-
ciency leads to a new state value denoted entropy. Like all state values, entropy within
a system must be constant in case a system is in a so-called steady state, i.e. when
none of the state values change in time. Entropy can be transported convectively by
the flow, it can be transported by heat crossing the system boundary, and it can be gen-
erated internally by dissipation, i.e. internal friction. Internal friction or dissipation
is a measure of the irreversibility of a process. The second law of thermodynamics,
which covers the balancing of entropy, finally answers the question of why cooling
is necessary in a thermal power plant. Every so-called heat engine, for example a
power plant, converts heat into mechanical/electrical energy, while another part is
released unused into the environment as heat. Once this mechanism is understood,
exergy is introduced as the part of the total energy that can be converted into any
other form of energy, e.g. electricity. The temperature of the combustion process
on the one hand and the ambient temperature on the other determine the maximum
efficiency. In other words, the temperature difference between which the heat engine
operates is therefore crucial for maximum efficiency.

Cyclic processes

However, the examples of a coal-fired power plant or an internal combustion engine


show an important field of work for the thermodynamicist: Heat is converted into
mechanical/electrical work to operate a machine. These processes are called clock-
wise cycles or thermal engines. Such a process is illustrated in Fig. 1.1, which shows
a Stirling machine. In contrast to clockwise cycles, a typical counterclockwise cycle
is a refrigerator or heat pump that utilises ambient heat at low temperatures to heat
a building on a higher temperature level.

1.2 Classification of Thermodynamics

The common thermodynamics textbooks are usually divided into three categories,
which are briefly presented here. Students of conventional mechanical engineering
learn the basics of technical thermodynamics and, depending on the subject, chem-
6 1 Introduction

ical thermodynamics. There are a number of textbooks for both categories, both in
German and English. Thermodynamics is described phenomenologically in these
fields. However, students of natural sciences deal with statistical thermodynamics or
statistical mechanics, which is addressed superficially in this book.

1.2.1 Technical Thermodynamics

This area of thermodynamics focuses on engineering tasks, i.e. the conversion and
utilisation of energy in thermal machines. This book introduces thermal machines
such as internal combustion engines and gas turbines, which are characterised by
the conversion of thermal energy into mechanical energy. Such machines work as
clockwise thermodynamic cycles. Refrigerators and heat pumps, however, are so-
called counterclockwise cycles.

1.2.2 Statistical Thermodynamics

Statistical thermodynamics investigates the properties of fluids, i.e. atoms or


molecules. Macroscopic state variables (pressure, temperature, entropy, enthalpy,
energy) are calculated by means of statistical mechanics. Many findings of this statis-
tical approach form the basis for the technical thermodynamics. In Part I, the focus is
on the so-called ideal gas and the corresponding gas equation, which can be derived
primarily from statistical thermodynamics. According to Boltzmann, entropy S is
proportional to the logarithm of the number  of microstates of a system. Microstates
refer to the ways in which the molecules or atoms of a trapped fluid can be arranged:

S = kB ln () (1.1)

The constant kB is the Boltzmann constant. However, the statistical thermodynamics


is not part of this book. For further details reference is made to Bošnjaković and
Knoche, see [11].

1.2.3 Chemical Thermodynamics

Chemical thermodynamics describes chemical and physical/chemical processes. Of


interest are, for example, the degree or limitations of conversion, i.e. the dynamic
chemical equilibrium, and the speed of a reaction. A chemical reaction is a material
conversion of reactants into reaction products that have different material properties
compared to the reactants. A chemical reaction is characterised by heat and mass
transfer. In chemical thermodynamics, a distinction is made between homogeneous
1.2 Classification of Thermodynamics 7

Fig. 1.2 Distinction thermodynamics (left)/heat transfer (right)

and heterogeneous chemical reactions. A homogeneous equilibrium is reached when


the reaction products occur in only one phase state, e.g. gaseous, liquid or solid. A
heterogeneous equilibrium is reached when the reaction products do not only occur in
one phase, but e.g. in liquid and gaseous state. Typical combustion processes occur
in internal combustion engines, jet engines, boilers and rocket engines. Methods
of chemical thermodynamics are treated in Chaps. 23 and 24. Therefore, chemical
thermodynamics is partially covered in this book, see Part III.

1.3 Distinction Thermodynamics/Heat Transfer

In addition to thermodynamics, heat and mass transfer are also part of the study of
thermal engineering. This raises the question of how these two topics differ from
each other. The most important aspects are shown in Fig. 1.2 and are explained below.

1.3.1 Thermodynamics

Thermodynamics defines systems and their states, which can be specified by temper-
ature, pressure, internal energy and others. It investigates how the state of a system
can be changed by process variables, i.e. external effects or influences that may act
across the system boundary. These process variables are work and heat, i.e. different
forms of energy. Thus, a system that was initially in a state of equilibrium reaches a
new state of equilibrium after a period of time.
Furthermore, the example in Fig. 1.2 shows a so-called homogeneous system,
i.e. there are no gradients in any state values: The temperature, for example, is
uniformly distributed, even on the wall where the heat is supplied from outside. A
possible temperature gradient within the system for heat transfer is not needed in a
thermodynamic perspective.
Unfortunately, thermodynamics does not explain how a system can release heat or
how heat is supplied to a system. The physical transport mechanisms are not explicitly
8 1 Introduction

explained in detail. In particular, a temperature gradient is required to initiate heat


transfer. Strictly speaking, a homogeneous system is therefore not able to transfer
heat. Although energy conversion can be explained with the help of thermodynamics,
it is not possible to design, for example, a heat exchange in detail. Thermodynamics
teaches what influence heat has on a system, but does not explain how heat can be
supplied or released.

1.3.2 Heat Transfer

Thermodynamics explains the correlation between a thermal potential, i.e. a tem-


perature difference, and a resulting heat flux. It follows the physical principle of
cause and effect. Every temperature difference means an imbalance. Intensive state
variables, i.e. pressure and temperature, have in common that they cause a transient
balancing process.2 Thermodynamics teaches that this balancing process is accom-
panied by the generation of entropy—the greater the imbalance, the more entropy is
generated.
However, the field of heat transfer explains how heat transfer works in detail:
conduction heat transfer in systems at rest, convective heat transfer, which is related
to fluid dynamics, and radiation, which works even in vacuum. Unlike thermody-
namics, which deals with homogeneous systems, heat transfer deals with spatial and
temporal temperature profiles. From a thermodynamic point of view, these systems
are heterogeneous, i.e. non-uniform. A typical non-uniform temperature profile is
shown in Fig. 1.2. These temperature profiles are essential for heat transfer and are
not covered by thermodynamics.

1.4 History of Thermodynamics

1.4.1 The Caloric Theory Around 1780

The french chemist Lavoisier 3 postulated the so-called Caloric theory around
1783. Hence, he assumed that heat is an invisible and weightless substance, called
“caloric”, which is located in the material between the molecules/atoms. Accord-
ing to Lavoisier’s approach, caloric can be transferred from one body to a second
body, which expands and warms up. The transfer is initiated due to an internal repul-

2 Everyone can observe this process when a hot cup of coffee is placed in the environment. An
imbalance occurs and the system coffee brings itself into a new state of equilibrium with its envi-
ronment. After a while, the coffee and the environment are in thermal equilibrium—identifiable by
a cold cup of coffee at ambient temperature. This is achieved by heat release from the coffee to the
environment.
3 Antoine Laurent de Lavoisier (26 August 1743 in Paris, 8 May 1794 in Paris).
1.4 History of Thermodynamics 9

sive force of caloric. Lavoisier assumed that caloric, like the elements, can not be
destroyed or generated. Thus, the amount of caloric remains constant. Lavoisier was
able to explain the effects of phase change with his theory: Supply of caloric causes
a phase change from solid to liquid. The reversal change of state, i.e. from liquid to
solid state, can be achieved by release of caloric.
The theory is, from today’s perspective, outdated as clarified by the following
simple example. Rubbing two bodies against each other, “generates heat” at the con-
tact point of the bodies. This frictional process can be sustained over a long period of
time without exhausting a finite reservoir of caloric. Thus, this is a contradiction to
Lavoisier’s theory. The process can rather be described with modern thermodynamics
in the form of the first law of thermodynamics: According to the first law, frictional
work causes an increase of the bodies’ internal energy, that results in a rising tem-
perature.4 However, heat is a process value,5 that only occurs when two bodies in
contact have different temperatures. An exchange of thermal energy, i.e. heat, takes
place between the two bodies until a thermal equilibrium is reached between the bod-
ies. In this thermal equilibrium, both bodies have the same temperature. However,
the phenomena, that heat supply respectively release can cause a phase change, still
corresponds to the caloric theory.

1.4.2 Thermodynamics as from the 18th Century

In order to show the history of modern thermodynamics, important scientists in the


field of thermodynamics are introduced below. Their contribution to thermodynamics
is briefly presented first. However, their theories are explained in detail later in this
book.
The french engineer officer N.L.S. Carnot 6 is considered to be one of the founders
of modern thermodynamics.
• He postulated the first modern theory of how work can be obtained from heat.
Today, this principle is realised in every modern thermal engine. A thermal engine
converts thermal energy, e.g. released by burning fossil fuels, into mechanical
work.
• Thermodynamic cycles are important in modern technology. They are charac-
terised by a working fluid that undergoes several changes of state until it finally
reaches its initial state. Cycles can, as already shown, be designed for different pur-

4 Mechanical energy is converted into dissipation/frictional energy, that increases the body’s internal
energy.
5 In thermodynamics, it is reasonable to distinguish between state values, i.e. parameters that

describe the internal state of a system, and process values that act on a system and can thus change
the state of the system.
6 Nicolas Léonard Sadi Carnot (1 June 1796 in Paris, 24 August 1832 in Paris).
10 1 Introduction

poses. For example, in thermal engines, see above, heat is converted into work.7
At a later stage, the thermal engine is referred to as a clockwise cycle. Carnot
raised the question of how this cycle could be optimised to achieve the maximum
amount of work. Carnot postulated that the thermal engine must avoid any internal
friction—in thermodynamics the word dissipation is used to quantify internal fric-
tion. Additionally, he found how heat, i.e. the driver of a clockwise cycle, needs
be supplied. Carnot derived a correlation for the maximum efficiency of reversible
cycles, i.e. cycles free of dissipation. This phenomena is discussed in detail when
the second law of thermodynamics is introduced.
The British private scholar J.P. Joule8 from Manchester and the German physician
J.R. Mayer 9 from Heilbronn refined the theory of heat.
• According to the principle of the equivalence of heat and work, heat is no longer
an indestructible substance, as was assumed for the caloric, see Sect. 1.4.1. Rather,
it is possible to convert work into heat and vice versa. This is an important finding
in thermodynamics. Heat and work influence the state of a system in the same
matter, they are convertible into each other. In the 19th century the principle of
equivalence replaced the caloric theory. This principle eventually led to the law of
conservation of energy, also known as the first law of thermodynamics.
• Joule proved the so-call “mechanical equivalence of heat” with a simple experi-
ment. This experiment is known as Joule’s paddlewheel and is explained in Chap. 2:
By supplying mechanical work to an adiabatic, i.e. perfectly insulated, vessel filled
with water, a rise in temperature can be observed.
• In addition, Joule showed that electrical energy can be converted into heat—
denoted as Joule heating. The amount of heat is proportional to the electrical
resistance and proportional to the square of the current.
The German professor R.J.E. Clausius,10 Professor of Physics (Germany and
Switzerland) has contributed significantly to modern thermodynamics and will also
guide us through this book.
• He provided the mathematical formulation of the two laws of thermodynamics.
The first law of thermodynamics predicts how process values, i.e. heat and work,
passing the boundary of the system affect the internal state of a system.
• Thus, Clausius introduced the state value internal energy, which can vary due to
external influences.
• The first law of thermodynamics can be applied to explain the theory of energy
conversion, but it does not take into account the limitations of energy conversion.
Consequently, according to the first law, it would be possible to design a machine,

7 In this case, heat is the expense, as we have to pay for the fuel needed to produce the heat. Work
is the benefit, as the thermal engine is designed to produce mechanical or electrical work.
8 James Prescott Joule (24 December 1818 in Salford near Manchester, 11 October 1889 in Sale

(Greater Manchester)).
9 Julius Robert von Mayer (25 November 1814 in Heilbronn, 20 March 1878 in Heilbronn).
10 Rudolf Julius Emanuel Clausius (2 January 1822 in Koeslin, 24 August 1888 in Bonn).
1.4 History of Thermodynamics 11

that permanently removes heat from a hot reservoir and converts it completely into
mechanical work, e.g. a coal fired power plant without any cooling tower. Since
there is no such power plant, cooling is required as part of the energy conversion:
However, any heat released to the environment reduces the electrical power. This
universal principle, that every thermal machine in which heat is converted into
work requires cooling for steady-state operation, is introduced as the second law
of thermodynamics. The physical explanation was given by Clausius and is closely
linked to the concept of entropy.
• For a quantitative formulation of the convertibility of energy, known as the second
law of thermodynamics, Clausius introduced a new state value “equivalent value of
transformation”. This equivalent value is denoted entropy, which is of fundamental
importance in thermodynamics. There is no conservation law for the state value
entropy as there is for energy, mass and momentum. Entropy can nevertheless
be balanced: It can be stored in systems, generated internally by dissipation, and
transferred by convective flow or heat exchange. Chap. 13 shows how the balancing
is done in detail. Entropy gives the changeability of energy a mathematical basis.
Furthermore, it explains why e.g. heat always follows a temperature gradient.11
Another important thermodynamicist of modern science was the Scotsman W.
Thomson,12 who was a professor of natural philosophy and theoretical physics in
Glasgow.
• Thomson gave an alternative formulation of the second law of thermodynam-
ics, also known as the theorem of dissipation of mechanical energy. He realised,
that any irreversibility inside a system, i.e mechanical friction or transient bal-
ancing processes, are associated with the generation of entropy. Every generation
of entropy in a system reduces its working ability. Thus, a new term exergy, i.e.
working capability, is introduced, which is closely linked to entropy. Hence, the
loss of exergy is a function of entropy generation, see Chap. 16.
• Entropy can, as already mentioned, be stored, transferred by heat or transported by
mass flows and generated by irreversibilities. Since entropy cannot be destroyed,
but can only be released by the system to the environment, e.g. by releasing heat
into the environment, the global amount of entropy is constantly increasing.
• Entropy S, absolute temperature T and heat Q are fundamental parameters in
thermodynamics and correlate as follows:

δ Q rev = T dS (1.2)

This is one of the essential equations explored in this book.

11 Think of a hot plate that is touched by your hand. According to the first law, it would be possible
for your hand to cool down while the temperature of the hot plate continues to rise. It is the second
law of thermodynamics that determines the direction of heat transfer. Do not touch a hot plate!
12 William Thomson, 1. Baron Kelvin, mostly as Lord Kelvin or Kelvin of Largs (26 June 1824

in Belfast, Northern Ireland, 17 December 1907 in Netherhall near Largs, Scotland).
12 1 Introduction

Fig. 1.3 Working principle of a steam machine

In addition to the fundamental principles in the 18th/19th century, the first tech-
nical applications were also developed at this time. First and foremost, the British
inventor T. Newcomen13 should be mentioned here.
• Newcomen has already designed in 1712 the first atmospheric steam machine, that
was applied in the mining industry in order to pump out the intruding water in the
tunnels. His invention and the application of the steam machine is strongly related
with the industrial revolution. The industrial revolution accelerated the technical
progress as well as the social development and was originated in England.
• The steam machine can be regarded as the first thermal engine, that converts
supplied heat into work. Fig. 1.3 shows the working principle of a steam machine.
The supplied heat vaporises the working fluid in a boiler, which is not shown in
Fig. 1.3. The vapour is subsequently lead into the valve chest, which can direct
the steam into a cylinder at two points. This sets both a piston and a piston rod
in motion, so that the flywheel starts to turn. By switching the steam inlet, the
piston can be set into axial oscillation so that the flywheel continues to rotate. The
switching of the steam inlet is achieved by a rotary valve, that is controlled by
the flywheel. Finally, the thermal energy is converted into mechanical, rotational
work. The efficiency of Newcomen’s machine was less than 0.5%, i.e. less than
0.5% of the supplied thermal energy of the fuel is converted into mechanical work.
• J. Watt 14 performed several technical improvements to the machine in 1769, that
lead to the English patent Nr. 913. The thermal efficiency of Watt’s machine was
up to 3%.

13 Thomas Newcomen (26 February 1663 in Dartmouth, 5 August 1729 in London).


14 James Watt (30 January 1736 in Greenock, 25 August 1819 in Heathfield, Staffordshire).
1.4 History of Thermodynamics 13

1.4.3 Thermodynamics in the 21st Century

Even though one might get the impression that thermodynamics is old-fashioned,
its philosophy and theoretical foundations are still relevant and applicable in the
21st century. Nowadays, thermodynamics is still an important subject needed for the
development of modern and innovative machines. This is the reason why students of
mechanical engineering still need profound knowledge in this field. The following
section lists current topics that have been developed on the basis of thermodynamic
principles. Most of these applications have been optimised over the years.

Combustion engine

The internal combustion engine is a thermal engine with internal combustion, i.e. a
fuel-air mixture is ignited in a cylinder. The chemical reaction releases heat. How-
ever, it differs from gas or steam turbines because internal combustion engines actu-
ally work discontinuously. Strictly speaking, even the Stirling engine presented in
Sect. 17.2.5 is not an internal combustion engine, as the combustion is external. Over
the years, the internal combustion engine has undergone constant development, which
is evident in falling fuel consumption with increasing vehicle mass. Another chal-
lenge is the increasingly stringent emission limits. According to the second law of
thermodynamics, the engine must transfer heat to a cold reservoir in order to operate
continuously. As will be shown later, heat release should occur at ambient tempera-
ture to minimise exergy loss. This maximises the temperature potential, which is the
driving force of the process. Nevertheless, the exhaust gas still contains exergy, e.g.
thermal exergy because the exhaust gas has a higher temperature than the environ-
ment, and kinetic energy due to the velocity of the exhaust gas. While in the past this
exergy leaving the vehicle was not recovered, current developments try to recover
the exergy from the exhaust gas. Two approaches are presented in Sect. 1.4.4.

Gas and steam turbines

A combined gas/steam turbine plant is another example for a thermal engine. The
exhaust heat of the gas turbine is utilised to operate a steam turbine process. Gener-
ally speaking, in a turbine a fluid expands from a high pressure to a lower pressure.
Thus, thermal energy is converted into mechanical energy. Similar to other thermal
engines, the thermal efficiency increases with the temperature at which heat is sup-
plied. Obviously, while optimising a thermal engine other disciplines, e.g. material
science, are required as well in order to design materials, that can withstand the
increasing temperatures, cf. Fig. 1.4.
14 1 Introduction

Fig. 1.4 Gas turbine (Schematic working principle)

Fossil and nuclear power plants

Basically, what has been mentioned for gas and steam turbines applies for power
plants as well: Continuously working thermal engine, working in-between two tem-
perature levels, have a theoretically maximum thermal efficiency, that depends solely
on the maximum (hot reservoir) and minimum (cold reservoir) temperature—called
Carnot efficiency. In the 2000s approximately 44% of the electricity produced in
Germany comes from firing hard coal and lignite. The averaged thermal efficiency
of all installed power plants was approximately 38%, compared with Watt’s steam
machine a considerable increase. Figure 1.5 shows the schematic working principle
of a fossil power plant as well as a visualisation in a p, v-diagram. The p, v-diagram
is an important tool in thermodynamics. In this case, it explains why a thermal engine
is supposed to be a clockwise cycle.

Compressors, pumps, blowers

These components are important parts needed for the design of thermodynamic
cycles. The task of an engineer is to develop components that are as efficient as
possible and which, theoretically idealised, are free of any irreversibilities. If these
theoretically friction-free components are also adiabatic, i.e. perfectly insulated, the
thermodynamicist names these components isentropic.
1.4 History of Thermodynamics 15

Fig. 1.5 Fossil power plant (left: Schematic working principle, right: Visualisation in a p, v-
diagram, liquid water is supposed approximately to be incompressible)

Propulsion systems for aircrafts and rockets

Aircraft turbines follow the same principle as gas turbines. In a first step, air is
compressed and fed into a combustion chamber. After the ignition of an air/fuel
mixture, thermal energy is released at high pressure. Within the turbine the exhaust
gas expands and thermal energy is converted in mechanical energy. The released
energy can be partially used to operate the compressor. In a finally step, the exhaust
gas leaves the nozzle and provides the aircraft’s propulsion. For supersonic flights
the nozzle needs to have a specific shape, called Laval nozzle, that will be introduced
in Part II.

Combustion systems

The areas of thermodynamics already presented clearly show that many of them
require a combustion process. For this reason, there is an entire chapter in this book
on the theory of combustion. When planning a combustion process, it is necessary to
understand the ratio in which fuel (e.g. fossil fuels such as coal, gas or oil) and air must
be mixed in order to achieve the desired combustion temperature and to guarantee
the emission limits. Although this technology could be considered old-fashioned, as
nowadays e.g. combustion engines are easily replaced by new, emission-free systems,
new systems, e.g. fuel cells, are also based on reactive processes.

Cryogenic systems

This area of thermodynamics deals with processes at very low temperatures. These
processes are essential for both gas separation and liquefaction. One well-known
16 1 Introduction

technical application is the Linde process,15 that utilises the Joule-Thomson effect.
While the temperature remains constant during adiabatic throttling16 of an ideal gas,
real gases are subject to the Joule-Thomson effect, which may lead to a decrease in
temperature. A detailed explanation of this effect can be found in Sect. 18.5.

Heating, ventilation and air conditioning (HVAC)

Air conditioning in buildings also requires knowledge of thermodynamics. In addi-


tion to temperature, which can be influenced by air conditioning, humidity is another
comfort criterion. Thermodynamic principles are used for the energy dimensioning
of HVAC systems. In particular, in Chap. 20 we will focus on gas mixtures and humid
air.

Compressions-/ and Sorption cooling machines

Cooling machines work according to a counterclockwise cycle that must be supplied


with mechanical/electrical energy. In the process, heat is taken from a cold reservoir
and finally transferred to a hot reservoir at a higher temperature. Figure 1.6 illustrates
the functional schematic of a cooling machine and the required components. In
contrast to clockwise cycles, that are quantified with a thermal efficiency, a so-called
coefficient of performance (COP) is defined for counterclockwise cycles. It is the
ratio of benefit and effort. For a cooling machine the benefit is the supplied heat at
low temperature, whereas the effort is the supplied mechanical/electrical energy of
the required compressor. In cooling machines that work according to the absorption
principle, the mechanically operated compressor is replaced by a so-called thermal
compressor.

Heat pumps

Heat pumps do have the same technical layout as cooling machines, including the
required components. However, the benefit differs: Heat pumps are commonly used
for heating purpose, thus the benefit is the heat released at high temperature. Similar
to the cooling machine, heat is taken from a cold reservoir, that might be ambient air,
soil or even water. Since it is a contradiction to the second law of thermodynamics, i.e.
heat follows the temperature gradient and does not move from cold to hot, mechanical
work is required to “pump” heat from a cold reservoir into a hot reservoir. Once again
the coefficient of performance characterises the efficiency this machine. Typically, the
COP is greater than one. If the COP is equal to one, the machine is an instantaneous

15 Named after the German engineer Carl von Linde (26 June 1824 in Berndorf, 16 November
1934 in Munich).
16 An adiabatic throttling is treated as isenthalpic—in case potential and kinetic energies are ignored!
1.4 History of Thermodynamics 17

Fig. 1.6 Cooling machine (left: Schematic working principle, right: Illustration in a T, s-diagram,
the compressor is supposed to be isentropic)

water heater. The entire heating purpose is covered electrically and thus is not the
preferable solution. A COP e.g. of six means, that for a specific heating requirement,
i.e. the benefit, of 6 kW a mechanical/electrical power of 1 kW needs to be supplied.
The rest of 5 kW is taken from the environment and does not constitute an effort.

Fuel cell systems

Conventional energy conversion systems are based, as described above, on com-


bustion processes. The heat being released at the combustion can be converted in a
thermal engine in order to generate mechanical/electrical work. However, in a fuel
cell a so-called cold combustion, i.e. a chemical reaction of hydrogen and oxygen,
occurs. Chemical bonded energy is directly converted into electrical energy. As the
only reaction product is water, this energy conversion is emission-free. However,
the required hydrogen needs to be generated, e.g. with renewable energies. In times,
where politicians debate of banning vehicles with internal combustion engines, we
currently live in an era of major changes. Hence, fuel cell technologies are very likely
applied in the near future.

Solar systems, geothermal systems, wind turbines

The thermodynamic principle apply for renewable energy systems as well. A detailed
introduction to these systems is not given in this book but can be found for example
in [12].
18 1 Introduction

1.4.4 Modern Automotive Applications

In the late 2000s, the electrification of vehicles started in Europe. This is an important
step for the automotive industry, as the conventional combustion engines have to be
replaced step by step by new drive systems. One possibility of electric driving is the
use of lithium-ion batteries for charging and discharging electrical energy. Energy
recovery by a generator during braking is possible with this technology. However,
several technical options are also being developed from all-electric driving to hybrid
drive, i.e. the combination of electric motor and combustion engine.
Nevertheless, the optimisation of the combustion engine did not stop then. Major
steps of several percentage points in thermal efficiency are of course not to be
expected, but small steps are also acceptable. According to the second law of ther-
modynamics, a thermal engine must release heat. Since the temperature level of the
released heat is usually higher than ambient temperature, it still contains exergy.
From a thermodynamic point of view, it makes sense to focus on this exergy in order
to utilise it. Two possible systems for recovering the waste heat are briefly presented.

Thermoelectric generator

Thermoelectric generator and thermocouples for temperature measurement are based


on the same physical principle, known as Seebeck effect, which describes a ther-
moelectric potential: Thermoelectric materials, e.g. Bi2 Te3 , generate an electrical
potential once a temperature gradient is applied. This requires one side of a ceramic
material to be hot while the other side is cold. From a thermodynamic point of view,
the temperature gradient triggers a heat flux that is partially converted into electrical
energy. An electrical efficiency can be defined for the conversion rate of the thermo-
electric element. Currently, the efficiency is below 10%, depending on the type of
ceramic.
For a technical application in a vehicle, both a hot and a cold reservoir are needed.
Figure 1.7 illustrates a possible application in an automobile. A thermoelectric ele-

Fig. 1.7 Thermoelectric generator (TEG)


1.4 History of Thermodynamics 19

Fig. 1.8 Waste heat recovery

ment is mounted on the exhaust pipe. In this position, the temperature is above
the ambient temperature, so that a potential heat flux occurs. Depending on the
state of the vehicle, the exhaust pipe can be bypassed with a flap. There are several
alternatives for providing a cold reservoir: Generally speaking, with regard to the
maximum temperature spread, cooling with ambient air would be the best variant.
Unfortunately, the heat transfer characteristic in this case is rather poor and would
lead to an increased heat transfer area. This is in contradiction to a light and compact
design. The preferred option is to use the cooling circuit of the combustion engine.
The temperature is higher than ambient temperature, but still sufficient to create a
temperature gradient across the element. The use of cooling water as a cold reservoir
guarantees a compact design, as the heat transfer is more efficient than with air, for
example.
The recovered waste heat is available as electrical energy, which can be used to
operate electrical auxiliaries or to relieve the load on the alternator. This example
shows that technical improvements are still possible today. In order to optimise such
technical systems, detailed knowledge of the energy conversion is required.

Turbosteamer

The basic idea of the so-called turbo steamer, see [13], is the same as for the thermo-
electric generator: the waste heat is to be recovered in order to improve the overall
energy balance of the vehicle.
In this application, the waste heat is used to heat a fluid in a secondary cycle, as
shown in Fig. 1.8. Actually, this system is a combination of an internal combustion
engine and a steam-based thermal engine: the recovered waste heat vaporises a fluid
that has previously been pressurised.
20 1 Introduction

The steam, which is under high pressure when it leaves the heat exchanger, can
then expand in a turbine, for example. In this step, thermal energy is converted into
mechanical energy. Part of this mechanical energy can be used to cover the energy
consumption of the pump required to pressurise the liquid. Once the mechanical
energy is available, it can be used for boosting or converted into electrical energy in
a generator.
As indicated in Fig. 1.8, the thermodynamic process is designed as a cycle. There-
fore, another component is required to close the cycle. After leaving the expansion
device, the liquid must be condensed to reach the initial, liquid state. This is realised
by a second heat exchanger. At this point, the cycle is closed and can start again.
The entire process can be regarded as clockwise thermal engine. Thermal energy is
converted into mechanical/electrical energy, whereby cooling is still required accord-
ing to the second law of thermodynamics. This cooling requirement is the reason
why the thermal efficiency of the turbosteamer is always less than 100%. All thermal
engines have this in common.
Part I
Basics and Ideal Fluids
Chapter 2
Energy and Work

It is known from physics lessons that energy is conserved. This means that the amount
of energy is constant, i.e. it can neither be generated nor destroyed. Thus, the principle
of energy conservation reads as:
Theorem 2.1 In a fully-closed system, with any internal mechanical, thermal, elec-
trical or chemical processes, the total amount of energy remains constant.
However, before clarifying in Chap. 3 what a fully-closed system is, let us first
focus on the term energy. Energy can be transferred to other objects or even trans-
formed into other forms, as discussed in the previous introductory chapter. Any work
done on a mass increases its energy and enables it to perform work itself. Energy is
therefore defined as the ability of a mass to carry out work, see e.g. [14].

2.1 Mechanical Energy

First, let us concentrate on what is known from mechanics: work is based on an


interaction between a force F and a mass m. This force performs work W when it
displaces the mass by a distance s. Work is the scalar product of force and distance,
see Fig. 2.1:
W = F · s. (2.1)

The scalar product obeys


W = F · s · cos α (2.2)

with α being the angle between force and distance. Hence, in case force and distance
are perpendicular to each other, no work is performed.
As is known from physics, power is the time derivative of work, i.e. the work that
is transferred per unit of time
© Springer Nature Switzerland AG 2022 23
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_2
24 2 Energy and Work

Fig. 2.1 a Definition of work and b power

δW
P= . (2.3)
dt
The work can hence be calculated by integration over a certain period of time:

2
W = P (t) dt. (2.4)
1

According to this definition of work, see Eq. 2.1, correlations can be derived for any
kind of mechanical energy, see the examples in Sects. 2.1.1, 2.1.2, 2.1.3.

2.1.1 Kinetic Energy

The kinetic energy of a mass correlates with the amount of work, that is required to
accelerate a mass, for example, from a state of rest to a certain velocity. A vehicle at
rest, i.e. state (1), is accelerated to a velocity of c2 in state (2). As shown in Fig. 2.2,
the function c (t) may have a non-linear characteristic. However, the question is how
much work is required to change the state of the vehicle from (1) to (2).

Fig. 2.2 Kinetic energy


2.1 Mechanical Energy 25

During the time period dt a distance of ds is covered. Thereby work is performed


according to Eq. 2.1:
δW = F · ds. (2.5)

The force obeys


dc
F =m·a =m , (2.6)
dt
the differential distance reads as
ds = c dt. (2.7)

Hence it is
dc
δW = m c dt = mc dc. (2.8)
dt
As indicated in Fig. 2.2, the change in velocity dc is positive when accelerating and
negative when decelerating. From the vehicle’s point of view, the work involved can
therefore be positive (acceleration) or negative (deceleration). Finally, the integra-
tion1 yields the total amount of work that must be expended to change the velocity:

2 2
1  2 
W = δW = m c dc = m c2 − c12 . (2.9)
2
1 1

This work is stored in position (2) as kinetic energy:

1  2 
E kin,12 = W12 = m c2 − c12 (2.10)
2

To calculate the work, therefore, it is not the function c(t) that is needed, but the
initial state (1) and the final state (2). In order to assign the kinetic energy to a state,
the reference level from which the counting starts must be determined. This can be
the state of rest, for example.

2.1.2 Potential Energy

A mass can be lifted in a gravitational field. This requires mechanical work, which is
stored as energy in the mass. The new vertical rest position depends on the amount

1Assuming, that the vehicle mass stays constant. If the vehicle is equipped with an internal com-
bustion engine, it must not emit exhaust gases into the environment. Or, alternatively, it is an electric
vehicle.
26 2 Energy and Work

Fig. 2.3 Potential energy

of work supplied. While moving a mass by a distance of dz within a gravitational


field, see Fig. 2.3, work is supplied:

δW = F · dz. (2.11)

The force is given by


F =m·g (2.12)

so that the following is obtained

δW = m · g dz. (2.13)

By integrating, one finally finds the total work that must be supplied to raise the mass
by a distance of h:

2 2
W = δW = mg dz = mg (h 2 − h 1 ) = mgh. (2.14)
1 1

This work is stored in position (2) as potential energy within the mass

E pot,12 = W12 = mgh (2.15)

As shown in Fig. 2.3 and proven by Eq. 2.15, it is necessary to define and establish a
reference level from which to count the potential energy.

2.1.3 Spring Energy

Another example of mechanical energy is the work needed to compress a spring, see
Fig. 2.4. The force is a function of the distance the spring is compressed and follows
2.1 Mechanical Energy 27

Fig. 2.4 Spring energy

F (x) = kF · x. (2.16)

Note that when the spring expands, work is released, while when the spring com-
presses, work is supplied, i.e. depending on the coordinate x, the force can be positive
or negative. Let x start at the reference position where there is no tension in the spring.
In differential notation the work is

δW = F dx = kF x dx. (2.17)

In order to calculate the entire work for a compression by x, an integration is done

2 x
1
W12 = δW = kF x dx = kF x 2 . (2.18)
2
1 0

This work is stored in position (2) as spring energy:

1
E spr = W12 = kF x 2 (2.19)
2

2.2 Thermal Energy—Heat

The previous examples have shown the correlation between work and energy from a
mechanical point of view. However, energy is more than just the ability of providing
work.
Heat, for example, is another form of energy: systems with different temperatures
transfer heat Q, i.e. thermal energy, at their interface as soon as they are in contact.
A hot plate that is switched on transfers heat to your hand when touching it.
The amount of heat obviously depends on the temperature difference between
hand and hot plate and also on the duration of the contact. The greater the temperature
28 2 Energy and Work

Fig. 2.5 Heat as thermal energy

potential between hand and heating plate, the greater the thermal energy transferred.
Once there is no more temperature difference, the heat transfer stops and the two
systems reach thermal equilibrium. Apparently, heat is another form of energy since it
has the power to change the state of the system. In this case, for example, the structure
of the hand will change. Based on experience, it is known that heat transfer follows
a temperature gradient. Thus, it has never been observed that the hot plate becomes
even hotter through contact while the hand cools down. In case no gradient occurs,
i.e. the systems having identical temperatures, no thermal energy is transferred, see
Fig. 2.5. This phenomenon is later explained by the second law of thermodynamics.
Consequently, the definition of energy must be expanded:
Theorem 2.2 Energy is not only the ability to perform work, but also the ability of
heating up a mass.
In thermodynamics it is important to distinguish between mechanical work W
and thermal energy Q.

2.3 Chemical Energy

Fuels also contain energy that is chemically bonded to the fuel and can be released by
a chemical reaction. This requires oxygen and fuel as reactants. During combustion,
exhaust gas is produced as a reaction product and energy can be released2 in form
of heat. This chemically bonded energy is usually defined by a so-called specific
lower/higher heating value and is discussed in detail in Part III. During combustion,
this chemical energy is converted into thermal energy. However, a fuel cell also
releases electrical energy in addition to thermal energy.

2 In case the reaction runs exothermically.


2.4 Changeability of Energy 29

2.4 Changeability of Energy

From experience, it is known that it is possible to transform the form of energy.


A mass fixed at a height h above a reference level carries potential energy that is
converted into kinetic energy as soon as the fixation is released. It is even possible
to convert kinetic energy into thermal energy, e.g. when braking a vehicle from a
velocity c1 to a standstill. Friction on the brake discs dissipates energy and eventually
releases it as heat into the environment. This leads to an extension of Theorem 2.1:
Theorem 2.3 Energy can exist in a variety of forms, such as electrical, mechanical,
chemical and thermal energy. It can be converted from one form to another. However,
it can not be destroyed.
This approach is further explored based on Joule’s paddle wheel experiment in
Sect. 2.4.1.

2.4.1 Joule’s Paddle Wheel

In this section, the focus is on the changeability of energy and it is shown that it
is necessary to introduce a new form of energy: Internal energy, which describes
the energetic state of a system. As the previous section has shown, energy can take
different forms and it can be transferred between systems, but it cannot be destroyed
or created out of nothing—this is known as the principle of conservation of energy.
Joule proposed a simple experiment, briefly sketched in Fig. 2.6.
First, a heat-insulated vessel is filled with a liquid, e.g. water. An impeller is
immersed in the water and is connected to two masses via a spindle. The two masses
are fixed on the outside at a precisely defined height above the ground. The entire
system is supposed to be in a state of equilibrium (1). Due to gravity the masses start
falling down as soon as the fixing is released. Obviously, while the masses descend,
the system is in imbalance. The potential energy of the two masses is eventually

Fig. 2.6 Joule’s paddle wheel experiment


30 2 Energy and Work

Fig. 2.7 Joule’s paddle wheel experiment, state (1) and (2)

converted into kinetic energy and rotational energy of the impeller. The impeller
starts to rotate and the liquid that comes into contact with the impeller is accelerated
by the friction3 . Distant fluid elements are still in rest. As a consequence turbulent
eddies within the water are generated. After a while, the energy carried by the eddies
is dissipated and the whole system comes to rest. Hence, a new state of equilibrium
(2) is reached.
Imagine that an observer was not present during the change of state from (1) to
(2), he only perceives the two equilibrium states (1) and (2). The question he will
probably ask himself is where the potential energy of the two masses has gone, since
is well aware of the principle of conservation of energy. Assuming that the walls are
massless, the entire energy of the masses must have been transferred to the water. The
observer therefore proposes to determine the temperature of the water, i.e. a measure
of the energy, first in state (1) and finally in state (2), see Fig. 2.7.4 Supposing it
were possible to measure the temperature accurately, a rise in temperature could be
observed from state (1) to state (2).

2.4.2 Internal Energy

With Joule’s paddle wheel experiment, see Sect. 2.4.1, it was found that potential
energy was first converted into kinetic/rotational energy. Eventually, this energy
dissipates and leads to an increase in the temperature of the water. Since the vessel is
perfectly insulated, the entire former potential energy needs to be within the water,
see Theorem 2.3. The fluid’s energy is called internal energy U . Thus, in state (1) the

3 This is because the first layer of water atoms sticks to the impeller’s surface, known as Stoke’s
no-slip condition.
4 His idea would be, the larger the temperature of the fluid is, the larger its energy content is.
2.4 Changeability of Energy 31

water has an amount U1 of internal energy. In the new equilibrium state (2), the system
has a greater amount of internal energy, i.e. U2 > U1 . Obviously, the energetic state
is indirectly characterised by temperature T , that is a measure of internal energy. The
larger the temperature is, the more internal energy is present. In general, the internal
energy U consists of
• thermal internal energy (kinetic energy of the molecules/atoms)
• chemical internal energy (molecular bond energy)—is treated in Part III
• nuclear internal energy (forces within the atomic cores)—is not treated in this
book
Commonly, only the thermal internal energy is taken into account as long as no
chemical/nuclear reactions take place. According to kinetic theory, thermal motion
at the molecular/atomic level correlates with temperature. In Chap. 24, however, an
approach is introduced that also includes the chemical binding energy. Colloqui-
ally, the heat content of a body is characterised as internal energy.5 Internal energy
describes the state of a system, it is therefore a state value. As it is to be explained,
the internal energy can never be measured directly, like it is the case with pressure
or temperature. Consequently, a state equation is required to calculate the internal
energy of a system, see Chap. 12. It is shown that the internal energy of an ideal gas
depends solely on temperature:

T
U − U0 = m cv (T ) dT. (2.20)
T0

However, for real fluids, as discussed in Part II, the internal energy depends on two
state values, e.g. pressure and temperature. In analogy to kinetic or potential energy,
a reference level is needed to determine from where to start counting the internal
energy. In general, the reference level is arbitrary, but must not be changed when
calculating a change of state, as is done later. A reference level can be defined as
follows:
U0 (ϑ = 0◦ C, p = 1bar) = 0. (2.21)

Respectively, for ideal gases:

U0 (ϑ = 0◦ C) = 0. (2.22)

Problem 2.1 Two identical vertical pipes with a diameter of D=0.1m are connected
by a thin tube and a valve, see Fig. 2.8. One pipe is filled with water up to a height
of H =10m. Water has a density of ρ = 1000 mkg3 , the other tube is initially empty.

5 Later in the book it is clarified that heat is not a state value. This means that a body never contains
heat, but it does contain internal energy. Every body has a temperature, a parameter that characterises
the state. Thermodynamically, on the other hand, heat is a process value that passes through a body
due to a temperature difference.
32 2 Energy and Work

Fig. 2.8 Sketch to Problem 2.1

Fig. 2.9 Solution to Problem 2.1

Now the valve is opened. After some time, an equilibrium of the water quantity is
reached. By what absolute value does the potential energy of the water change?

Solution (Alternative 1)

Starting with state (1), each slice dz of the column of water possesses mass dm and
potential energy dE pot , see Fig. 2.9 and Eq. 2.15:
π 2
dm = ρ dV = ρ D dz (2.23)
4
respectively
dE pot = dm gz. (2.24)

Hence, it is
π 2
dE pot = gzρ D dz. (2.25)
4
To calculate the entire potential energy in state (1) Eq. 2.25 needs to be integrated, i.e.

H H
π 2 π H
E pot,1 = dE pot = gzρ D dz = ρ D 2 H g . (2.26)
4 4 2
0 0
2.4 Changeability of Energy 33

By analogy state (2) follows6

H/2 H/2
π π H
E pot,2 =2 dE pot = 2 gzρ D 2 dz = ρ D 2 H g . (2.27)
4 4 4
0 0

Finally, the change of potential energy yields

π 2 H
E pot,12 = E pot,2 − E pot,1 = −ρ D H g = −1926.2 J (2.28)
4 4

Solution (Alternative 2)

Instead of using a differential approach as in alternative 1, the entire mass can be


regarded as concentrated in the centre of gravity, illustrated in Fig. 2.9. Thus, state
(1) obeys
H
E pot,1 = mg . (2.29)
2
Respectively, state (2) follows

H
E pot,2 = mg . (2.30)
4
Thus, the change of potential energy yields

H
E pot,12 = E pot,2 − E pot,1 = −mg . (2.31)
4
Substituting the mass, see Eq. 2.23, leads to the same result as before in alternative
1, i.e.
π H
E pot,12 = − ρ D 2 H g = −1926.2 J (2.32)
   4
4
m

The negative sign indicates, that the potential energy from state (1) to state (2)
decreases.

6 The factor 2 takes into account that there is a left and right water column.
Chapter 3
System and State

This chapter clarifies what a thermodynamic system is and how its state can be
described. Every system is separated from an environment by a system boundary.
This is done analogously to other technical disciplines, e.g. in technical mechanics.
First, the permeability of the system boundary is categorised. After classifying the
system and the system boundary, the next step is to quantify the internal state of a
system. This leads to so-called state values, e.g. pressure and temperature, which
determine the state of a system. Our everyday experience shows that the state can be
changed by external influences across the boundary.1

3.1 System

What is a system?

The definition of a thermodynamic system (brief: system) is prerequisite for perform-


ing thermodynamic balances. Figure 3.1 shows the principle of a thermodynamic
system. It can consist of a single compound, a body or an arrangement of several
bodies, apparatuses or machines. A system can be at rest or even in motion compared
to its environment. As indicated in Fig. 3.1 it is characterised by a system bound-
ary that separates system and environment. The system boundary can be an actual
boundary, e.g. a cylinder wall, or it can even be virtual. Environment and system
can potentially influence each other, as mass, momentum and energy can be trans-
ferred across the system boundary. Mass, momentum and energy are conservation
variables, i.e. their absolute amount remains constant. However, non-conservation

1 The fluid inside a bottle can be heated, e.g. by a lighter. This changes the fluid’s inner state, which
is given by the temperature, for example. The lighter is therefore an external influence that acts
across the system boundary.
© Springer Nature Switzerland AG 2022 35
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_3
36 3 System and State

variables can also pass the system boundary and can be balanced.2 The greatest
difficulty in solving thermodynamic problems is to find a most appropriate system
boundary for the problem at hand. Although a system boundary itself can never be
wrong, an inappropriately chosen boundary can lead to many interfaces and cause
difficulties in taking all relevant influences into account. Therefore, this book dis-
cusses numerous examples that demonstrate the importance of defining a system
boundary. The following questions are addressed and answered:
• What is transferred across the system boundary? The specification of permeability
leads to an important classification of systems.
• How can the state of a system be described? State values are needed that charac-
terise the inner state.
• How can the internal state of the system be specifically influenced by external
impact? The answer to this question leads us to the first law of thermodynamics,
see Chap. 11.
Example 3.1 Figure 3.2 shows a simple example of a system, everyone knows from
everyday experience. Your bank account can be treated as a system. In this case the
amount of money M is the internal state (i) of such a system. The system boundary
(B) separates your account from any financial environment and makes a balancing
for your bank possible. As you know, your account is subject to several influences
across its boundary. Some of them increase the internal state, e.g. your salary or the
bank’s interest income. But there are also influences that have the opposite direction,
i.e. decrease the amount of money M, e.g. your monthly rent as well as leisure
activities. Obviously, there is no difficulty in balancing a bank account: What comes
in increases the balance, what goes out decreases the amount. Starting with the initial
money available M0 and the fluxes across the system boundary, it is easy to calculate
the money stock M at a later stage:
 
M = M0 + Mi,in − M j,out . (3.1)
i j

This balancing method can be carried out with any extensive state value3 Z . This type
of balancing is conducted in thermodynamics. If all state values Z , e.g. the quantity
of money M, do not change in time, the system has reached a state of equilibrium or
a stationary state:
dZ
= 0. (3.2)
dt
Returning to the previous example, the general balance equation is as follows for a
stationary case  
Mi,in = M j,out . (3.3)
i j

2 Even though they can be generated, e.g. entropy, or they can be destroyed, e.g. exergy.
3 The expression extensive state value is specified in the next chapters. Extensive state values have
in common that they can be counted.
3.1 System 37

Fig. 3.1 Definition of a system, (i) indicates the internal state, (B) the system boundary

Fig. 3.2 Example of a bank account

On the way to steady state, the system may be subject to various transient effects
indicated by transient states.

3.1.1 Classification of Systems

Heterogeneous systems

A heterogeneous system consists of several phases. A phase is a region in which


each physical property is uniform. In the example shown in Fig. 3.3, the system con-
38 3 System and State

Fig. 3.3 Heterogeneous


versus homogeneous systems

sists of cylinder, piston, gas and environment. Each component forms a phase from
a thermodynamic point of view. Wall and piston are in a solid state, while gas and
environment are assumed to be gaseous. If, for example, the density is measured in
the entire system, there is obviously no uniform distribution, but a step, i.e. a non-
uniformity, can be measured at each interface. Obviously, it is impossible to assign a
single density4 to such a heterogeneous system. In the given example, however, a het-
erogeneous system can be divided into several homogeneous subsystems. Figure 3.4
shows an example of heterogenous systems. In many chemical engineering processes
granular media needs a drying. This can be achieved by fluidising the media with
air. However, the more air is used, the more turbulent the two-phase flow becomes. It
is found that both heat and mass transfer improve. The same technique can be used
for the combustion of coal and sludge. The required air fluidises the two-phase flow
and leads to good heat transfer properties, see [15]. From a thermodynamic point
of view this two-phase flow represents a heterogeneous system. According to [16],
Fig. 3.5 shows a diesel injection spray at high pressure. This is another example of a
two-phase flow consisting of air as gaseous phase and diesel as liquid phase. Thus,
this is a heterogeneous system as well. The fuel jet profile produced by the injection
nozzle is crucial during direct diesel injection. In best case, the injection initialises
small droplets in the outer region and coarser droplets in the internal region of the
spray. The small droplets spark first when the combustion process starts. As a result,

4The density of a system is a state value like pressure, temperature and many others, see Sect. 3.2.
The assignment of state values to a heterogeneous system can lead to difficulties because they can
be ambiguous.
3.1 System 39

Fig. 3.4 Heterogeneous systems—fluidised bed, see [15]

the fuel burns comparatively slowly and evenly. Therefore, the pressure rise in the
combustion chamber is rather smooth.
As already mentioned, a system is heterogeneous, for example, if several states
of aggregation are present. A simple model to explain states of aggregation is shown
in Fig. 3.6. Three states of aggregation are possible:
• Solid: stable in shape and volume, large intermolecular forces, strong lattice among
neighbouring molecules
• Liquid: stable in volume, interaction among neighbouring molecules, distance
between molecules constant at first approach
• Gaseous: No bonding between the molecules, small intermolecular forces, elastic
impact between the molecules
From now on, when talking about fluids, a gas or a liquid is meant. In Part I, fluids
are idealised: Gases are assumed to behave like an ideal gas, i.e. the molecules
or atoms of the gas are treated punctiformly, interactions between neighbouring
particles are neglected. This leads to quite simple equations of state. Unfortunately,
40 3 System and State

Fig. 3.5 Heterogeneous systems—diesel spray formation, see [16]

Fig. 3.6 Aggregate states from left to right: solid/liquid/gas

not all phenomena can be explained with this simplified approach: The so-called
Joule-Thomson effect, which explains why gases change their temperature when
they are adiabatically throttled, is a well-known deviation from ideal gas behaviour.
However, this phenomenon is explained in Part II: Real gases are introduced and
3.1 System 41

allow the Joule-Thomson effect to be calculated. The physical description of the


gaseous state becomes more complex at this point than for ideal gases.
The liquid fluids in Part I also follow a simplified idea: they are treated as incom-
pressible, i.e. they retain their volume even if the pressure or temperature changes.
This characteristic is further investigated in Part II: even if the compressibility is
rather low, this effect is taken into account. In addition, changes in the state of aggre-
gation are thermodynamically investigated in Part II.

Homogeneous systems

Any heterogeneous system can be further subdivided as long as homogeneous sys-


tems occur. In contrast to heterogeneous systems, the properties within such a system
are uniform, see Fig. 3.3. Taking density as an example, it is homogeneously dis-
tributed within the selected subsystems. Consequently, the number of interfaces or
boundaries increases when a heterogeneous system is subdivided. The homogeneous
system gas has two environments, namely piston and wall, both of which influence
the system gas. When solving thermodynamic problems, however, it is advisable in
most cases to start with a homogeneous system. Every system is uniquely defined by
a system boundary, which can basically have any shape. Most difficulties in balancing
a system occur when the system boundary is not uniquely chosen. Both appropriate
sketches and experience are needed to avoid these difficulties. Later in this book,
several examples are presented.

3.1.2 Permeability of Systems—Open Versus Closed Systems

An important distinction between systems is made with respect to the permeability


of the system boundary for mass. Closed systems are characterised by no transfer of
mass across the system boundary, i.e. the mass within the system remains constant.
Open systems allow mass to pass their system boundary. However, the total mass
within such an open system can still be constant, e.g. in a steady state, so that the
incoming mass must be balanced by the outgoing mass. An overview of the different
systems is shown in Fig. 3.7:

Insulated or fully-closed system

An insulated or fully-closed system does not allow


• mass transfer m
• transfer of work W
• transfer of heat Q
42 3 System and State

(a) (b)

(c) (d)

Fig. 3.7 Permeability of systems

across the system boundary. Space, for example, is considered a fully-closed system
in which there are no fluxes at its boundaries.5 With regard to technical applications,
a closed, perfectly insulated thermos can be treated as a fully-closed system as long
as the thermos is at rest, i.e. it is not moved: No mass crosses the system boundary,
the insulation prevents heat leaving the system6 and no mechanical work acts on the
thermos as long as it is at rest. Fully-closed systems represent a perfect ideal state
which, however, can hardly be achieved in technical reality, see thermos.

Closed system

In a closed system, no mass transfer m across the system boundary is permissible.


However, the transfer of work W and heat Q can occur. A cylinder filled with an
ideal gas and sealed with a movable piston is treated as a closed system as long
as the sealing is completely gas-tight. By moving the piston, work can be supplied
respectively be released and the gas inside can additionally be heated or cooled.

5 Though nobody has ever visited the boundary of space...


6 Every student knows, that a perfectly insulated thermos does not exist. After a three-hour lecture
the coffee inside a thermos is inedible. However, in thermodynamics ideal systems are assumed to
exist.
3.1 System 43

Open system

In contrast to a closed system, in an open system mass m can cross the system
boundary. The transfer of work W and heat Q is also possible. Many technical
applications are treated as open systems: A fluid flow through a channel can be
treated as an open system since mass enters and leaves the channel and heat can also
be transferred depending on the environment. Turbines and compressors are also
open systems, but unlike a flow channel, additional work W is released/supplied.

Heat insulated or adiabatic system

Both open and closed systems can be adiabatic, i.e. no heat passes the system bound-
ary. Technically, there are two possibilities for adiabatic systems: First, the system
has perfect insulation. Secondly, the system and the environment have the same
temperature, so there is no driver for heat transfer, see Fig. 2.5.

3.1.3 Examples for Thermodynamic Systems

Cooling machine

Figure 3.8 outlines several ways of defining a system. The sketch on the left in Fig. 3.8
shows a so-called counterclockwise thermodynamic cycle, i.e. a cooling machine or
heat pump. Although such cycles will be discussed later, see Chap. 15, the example
is already suitable to discuss the topic of system boundaries:
A system boundary can be chosen such that all components are part of the system,
see Fig. 3.8a. This is denoted overall or integral balance. According to our recent
definitions, such a system is closed: No mass enters or leaves the system. Regarding
the permeability of the boundary, the sketch shows that a heat flux Q̇ 0 enters while
another heat flux Q̇ leaves the system. Furthermore technical power Pt is supplied
to the system. In Sect. 15.2.1 it is shown that a cooling machine, i.e. a refrigerator,
is supplied with a heat flow Q̇ 0 of a rather low temperature.7 The component that
absorbs the heat flux is called evaporator. Since energy cannot be destroyed, the
heat absorbed at the low temperature must be released again. This happens in the
condenser, which releases heat at a higher temperature level.8 Obviously, heat is

7 Just think of your refrigerator at home: A bottle of beer that needs to be chilled is placed in the
fridge. There is a low temperature in this compartment. At this low temperature, however, a flow
of heat must leave the beer bottle to reduce its internal energy, i.e. to cool it down. This heat is
supplied to the refrigerant cycle.
8 You may have noticed that the back of your refrigerator has a higher temperature than the sur-

roundings and therefore releases heat.


44 3 System and State

(a)

(b)

(c)
Kon

Fig. 3.8 Cooling machine on the left: thermodynamic cycle on the right: possible system boundaries

shifted from a lower temperature to a higher temperature. This is contrary to our


everyday experience: heat always follows the temperature gradient from hot to cold.9
In a cooling machine, a heat flux must be lifted from cold to warm, i.e. against
the natural gradient. This requires technical power Pt .10 In Sect. 15.2.1 the machine
is explained with a state variable entropy: In steady-state operation, the entropy
inside the machine, like all other state variables, must be constant in time. Heat that
enters the machine at a low temperature brings a lot of entropy11 with it. To keep the
entropy in the machine constant, a greater flow of heat must be released at a higher
temperature level. According to the first law of thermodynamics, the heat released
can be increased if additional power is supplied to the system. The reason for this is
shown as soon as entropy is introduced.
Theorem 3.1 For most thermodynamic problems it is wise to start with an integral
balance first. If this is not sufficient in order to solve the problem, one should think
of gathering additional information, e.g. by balancing further systems.
Applying this theorem, additional information can be obtained by choosing a
system boundary as shown in Fig. 3.8b: The boundary separates the fluid flow into
and out of the component condenser: The condenser is an open system as the mass
flux ṁ 1,in enters the system and ṁ 1,out leaves the system. If the system boundary
contains only the mass flux of the refrigerant, as in case (b), the heat flux Q̇ becomes
visible at the boundary. A second mass flow that absorbs the released heat is then
not part of the system boundary.

9 Also known as second law of thermodynamics. Placing your hand on a hotplate demonstrates the
second law of thermodynamics impressively.
10 This is the reason why the fridge needs to be connected with a plug.
11 Thus, entropy obviously has something to do with heat and temperature!
3.1 System 45

Fig. 3.9 Examples for thermodynamic open systems

Alternatively, the boundary can also include the entire condenser, as shown in
Fig. 3.8c. In the condenser, the working fluid must release heat, which is transferred
to a second mass flow.12 With this system boundary there are two mass fluxes entering,
namely ṁ 1,in and ṁ 2,in . Both mass fluxes need to leave the component in steady state.
Since the heat flux Q̇ occurs internally, i.e. is not visible at the system boundary, it
is not considered in (c) and does not directly affect the thermodynamic balances.
Theorem 3.2 It is essential to clearly identify the system boundary and the physical
quantities passing it.

Passive versus active systems

Several components that are needed in the course of this book are briefly introduced,
see Fig. 3.9. All of them are open systems, i.e. mass crosses the system boundary.
However, it is useful to distinguish between active and passive systems. Passive
systems are actually work-isolated systems, i.e. no work/force passes the system
boundary. Examples of passive systems are:
• Heat exchangers required to transfer heat from one mass flow to another. In doing
so, one flow is cooled while the other is heated.13

12 In case of a fridge the working fluid releases heat to the ambient air, that can be regarded as
second mass flow. This ambient air rises due to buoyancy effects.
13 At least as long as it is sensible heat, i.e. the fluids are not subject to a phase change. Part II deals

with fluids that can also change their state of aggregation. If this happens e.g. isobaric, the heat
46 3 System and State

Fig. 3.10 Overview state values

• A throttle valve reduces the pressure of a mass flow by internal friction, i.e. dissi-
pation.
• Tubes or channels that conduct a fluid flow.
In contrast to passive systems, an active system is characterised by work/force
passing the system boundary. Examples of active systems are:
• Turbines as thermal machines that release work/force across the system boundary
as the pressure of the fluid decreases.
• Compressors are thermal machines as well. Instead of reducing the pressure of the
fluid, the pressure increases as it flows through the machine. Therefore, to increase
the pressure, work must be supplied across the system boundary.
• The stirred vessel in Fig. 3.9 is an active system as long as the stirrer transfers work
across the interface. Two possibilities are conceivable: If electrical or mechanical
work is supplied to the stirrer from outside, work is fed into the system. The other
possibility is that the flow in the system causes the stirrer to rotate. This work can
be transferred mechanically to the environment, so that work is released from the
system. The shaft work that is the cause of the transfer of work through the stirrer
is examined in more detail in Sect. 9.2.7.

3.2 State of a System

Now that it is known how to classify systems and how to define system boundaries,
the next step is to describe the internal state of a system. The thermodynamic state
of a system is clearly quantified by so-called state values, see also Sect. 5.1. Con-
sequently, changes within a system can be easily detected and quantified by state
values. A general classification of state values is shown in Fig. 3.10. A more detailed
explanation follows in the next sections. Figure 3.11 shows exemplary state values
for a closed system. The following sections summarise the various categories.

supplied is not utilised to modify the temperature, but to carry out the phase change. The heat is
then called latent heat and the phase change is isothermal.
3.2 State of a System 47

Fig. 3.11 Overview state


values for a closed system

Table 3.1 Temperature measurement


Method Measured quantity
Liquid thermometer (mercury, alcohol) Thermal expansion of a fluid
Gas thermometer (He, Ar, N2 ) Thermal expansion of a gas
Thermocouple (e.g. Ni/Cr-Ni) Electrical voltage (Seebeck-effecta ) of a metal couple
Resistance thermometer (e.g. Pt 100) El. resistance of a metal wire
Radiation thermometer (pyrometer) Impact of the detected heat radiation (e.g. el. voltage
of a photo diode)
a The Seebeck effect is a thermoelectric effect. Any thermoelectric element produces a voltage

potential when there is a temperature difference between its two ends, i.e. a temperature potential

3.2.1 Thermal State Values

Thermal state variables have in common that they can be easily measured. Therefore,
the following variables can be applied to determine the state of a system:
• Pressure p, [bar, Pa]14
• Temperature ϑ, T , [◦ C, K]  
  m3
• Volume V , m3 , specific volume v = V
m
= ρ1 , kg

The volume of a system can be quantified by geometrical means. If the volume


of a system is divided by its mass, one gets the so-called specific volume, i.e. the
reciprocal of density.
Sensors are available in order to measure the pressure as well as the temperature
of a system. The temperature can be specified on a Celsius scale, i.e. [ϑ] = 1 ◦ C, for
instance or on a Kelvin scale, i.e. as absolute temperature [T ] = 1 K. Most thermo-
dynamic equations require the absolute temperature T , which is essential in Chap. 6.
How can temperature be measured?
There are several ways to determine the temperature of a system, see Table 3.1, which
also contains the underlying physical principle.

14Just keep in mind, that Pa is the SI-unit. Nevertheless, for many applications bar is the more
common unit. The conversion follows 1 × 105 Pa ≡ 1 bar.
48 3 System and State

Temperature scales
As already mentioned, there are several possible temperature scales. In thermody-
namics, however, the absolute temperature T is the most important. It is a base unit
like e.g. length, mass and time and is measured in

[T ] = 1 K. (3.4)

1 K is defined as the 273.16th part of the thermodynamic temperature of the triple


point of water.15 The triple point is the state in which a fluid in thermodynamic
equilibrium can be liquid, solid and vapour at the same time. For water this occurs
at T = 273.16 K and p ≈ 611 Pa. The absolute temperature T starts counting at
absolute zero T = 0 K. At this state any molecular motion freezes. However, the third
law of thermodynamics states that the absolute temperature can approach absolute
zero arbitrarily without reaching it. In this book, nevertheless, the Celsius scale is used
in addition, which, compared to the absolute temperature scale, is shifted according to

ϑ = T − 273.15 K. (3.5)

Be aware that temperature differences are always expressed in K, i.e.

T = ϑ2 − ϑ1 (3.6)

with
[T ] = 1 K. (3.7)

3.2.2 Caloric State Values

In contrast to thermal state values, caloric state values cannot be measured, i.e. there
is no sensor to quantify a caloric state value. The internal energy U has already been
introduced as a measure of the kinetic energy of fluctuating molecules or atoms. It is
needed to explain the Joule’s paddle-wheel experiment, see Sect. 2.4.1. This section
has shown that the internal energy in ideal gases or ideal liquids correlates with the
temperature. However, there is no sensor with which one can determine the internal
energy directly. Consequently, an equation of state is required to calulate the internal
energy.
Although other caloric state values16 have not yet been introduced, there are two
others besides internal energy U that are of particular importance in thermodynamics:
• Internal energy U , [J]
• Enthalpy H , [J]

15 The thermodynamic explanation of the triple point is given in Part II!


16 Enthalpy and entropy are introduced at this point, although the physical explanation will follow
at a later stage. At present, it is simply assumed that they exist.
3.2 State of a System 49
J
• Entropy S, K
As already mentioned, it is not possible to measure these caloric state values. There-
fore, in Chap. 12, equations of state are derived to calculate caloric state values. A
general approach is followed, which is simplified for ideal gases and incompressible
liquids. Ideal gases and incompressible liquids are treated in Part I. Real fluids, which
may even undergo a phase change, are dealt with in Part II.

3.2.3 Outer State Values

In addition to the inner state, which is indicated by thermal or caloric state variables,
the so-called outer state values also quantify a system: The outer state refers to the
centre of gravity of a system, which can be at rest or in motion. With respect to
a reference level, a system possesses both potential energy, given by a coordinate
z perpendicular in the gravitational field, and kinetic energy, which correlates with
its velocity c. Both kinetic and potential energy have already been introduced in
Sects. 2.1.1 and 2.1.2.

3.2.4 Size of a System

So far, the internal state of a system has been introduced. Pressure and temperature,
for example, do not contain any information about the size of a system. To specify
the size of a system, extensive state values are required. Extensive state values are
part of Sect. 3.2.5: the size of a thermodynamic system can be quantified by its mass
m, its volume V or also by its molar quantity n.

3.2.5 Extensive, Intensive and Specific State Values

Another important distinction in thermodynamics is made between intensive and


extensive state values. The basic idea is shown in Fig. 3.12.
Intensive state values
With respect to Fig. 3.12, the intensive state values do not change when a system is
divided into subsystems. They are independent of the mass of the system:
• Pressure p
• Temperature T
Imagine a cup of coffee being decanted into two smaller cups. This has no influence
on the pressure, which is only imposed by the environment. If the process of decanting
is neglected, the temperature of the coffee does not change either. This behaviour is
characteristic of intensive state values.
50 3 System and State

Fig. 3.12 Extensive and


intensive state values

Extensive state values


In contrast to intensive state values, extensive state values Z change their magnitude
by dividing the system or by merging subsystems in proportion to the mass. In case
of the example documented in Fig. 3.12, the total volume V can be calculated by
adding up the volumes of the two subsystems

V = V1 + V2 . (3.8)

Extensive state values Z are


• Volume V
• Internal energy U
• Enthalpy H
• Entropy S
Extensive state values Z can be balanced, cf. Eq. 3.9, even if they do not satisfy
the quality of a conservation variable, i.e. even if there is a source, e.g. entropy
generation, or a sink,17 e.g. exergy loss:

dZ  
= Ż in − Ż out + Ż Source . (3.9)
dt
According to Fig. 3.13, the quantity of an extensive state value Z within a system
changes with time due to an incoming flux that increases the state value within the
system and an outgoing flux that decreases the quantity within the system. Both
sources and sinks also have an influence on the temporal variation of Z .18 Mind, that

17A sink can be treated by Ż Source < 0.


18Just think of your bathtub: the amount of water as a function of time in the tub is influenced
by both the flux of incoming water and the mass flux at the outflow. This example is described by
Eq. 3.9, although this example does not contain any sources or sinks. Another example has already
been given in Fig. 3.2.
3.2 State of a System 51

Fig. 3.13 Extensive state


value

Ż represents the flux of Z , i.e. the quantity of Z per second. A system in steady state
is characterised by no temporal change of any state value inside the system:

dZ  
=0= Ż in − Ż out + Ż Source . (3.10)
dt
This can be given in another notation
 
Ż in + Ż Source = Ż out . (3.11)

In other words: The inflow of an extensive state value equals the fluxes out of the
system!

In case the temporal development of a state variable is of no interest but only the
overall change counts, the differential Eq. 3.9 simplifies to a difference equation
 
Z = Z in − Z out + Z Source . (3.12)

Specific state values


The dependency of extensive state values Z on the mass of the system can easily be
eliminated by referring Z to the mass of the system m, i.e.:

Z
z= (3.13)
m
respectively

z= . (3.14)

z then is called specific state value. Examples for specific state values are:
 3
• Specific volume v = mV = ρ1 , mkg
 
• Specific internal energy u = Um , kg kJ
 
• Specific enthalpy h = mH , kg kJ
 
• Specific entropy s = mS , kgkJK
52 3 System and State

Fig. 3.14 Intensive versus


specific state value according
to [1] Left: thermodynamic
equilibrium possible, though
ρ1 < ρ2 Right:
thermodynamic equilibrium
impossible, if p1 = p2

Specific state values remain constant when a system is divided19 or merged from
several identical systems. Specific state values thus behave like intensive state values.
In contrast to intensive state values, specific state values are not driving forces for
state changes, see Fig. 3.14. The picture on the left in Fig. 3.14 shows a liquid and
a vapour which both have the same pressure p and the same temperature T , i.e.
they are in thermodynamic equilibrium. Systems like this are investigated in Part
II: Real fluids can undergo a change of aggregate state. Just think of water that is
heated and eventually vaporises. During vaporisation, liquid and vapour can occur
at the same time in thermodynamic equilibrium. In Part II the term wet steam is
introduced to describe this phenomenon. Referring to Fig. 3.14, initially both sub-
systems are separated by a dividing wall. Once the wall is removed, the system does
not drive itself into a new thermodynamic equilibrium, though the densities ρ1 = v11
and ρ2 are different. Obviously, the specific state value v is not able to trigger a new
thermodynamic state. One could also reverse the initial stratification, i.e. the liquid
is at the top, the vapour at the bottom, but even in this case there is no mixing, i.e.
formation of a homogeneous system. The densities of liquid and vapour will not
equalise. Only gravity would ensure that a stable stratification is created over time,
i.e. liquid at the bottom, vapour at the top.
The focus is now on the right picture in Fig. 3.14. In this case, two different,
but ideal gases are separated by a dividing wall. Initially, both of them have the
same temperature T and identical densities ρ = ρ1 = ρ2 . However, the pressures are
different, i.e. p1 = p2 . In Chap. 6 it is clarified how pressure, density and temperature
are related and how it is possible to achieve an arrangement as shown in Fig. 3.14.20
According to the thermal equation of state, for constant ρ and T , the following
results: p1
= ρT (3.15)
R1
p2
= ρT (3.16)
R2

p1 R1
= . (3.17)
p2 R2

19 To come back to the example with the cup of coffee: If the coffee is decanted into two cups, the
mass m and the volume V change, but the density ρ = 1v = m V remains constant.
20 However, the equation is given at this time, though the explanation follows later on.
3.2 State of a System 53

Obviously, it is the difference of the gas constants that makes the state in the right
picture of Fig. 3.14 possible. However, if the wall is removed, the state of the system
changes according to our everyday experience. This is due to the potential of the
intensive state value p, which forces the system into a state of equilibrium. Pressure
equalisation occurs, the system strives towards a new equilibrium in which the gases
are finally perfectly mixed by diffusion processes. The system is homogeneous.
Mixing processes of ideal gases are examined in Chap. 19.
Theorem 3.3 Every change of a thermodynamic equilibrium state is based on a
change of an intensive state value. A change of extensive state values does not cause
a change of a thermodynamic equilibrium. Specific state values behave as inten-
sive state values, though they do not have the power to drive a system in a new
thermodynamic equilibrium!
Molar state values
Instead of using the mass of a system m to calculate specific state variables z, one
can alternatively use the molar quantity n. By this method, one obtains the molar
state values z M :
Z
zM = . (3.18)
n
The definition of molar quantity n is rather simple and related to the following
question: How many atoms of carbon isotope21 12C are required in order to get
12 g of 12C?22 As known from chemistry, 6.022045 ± (0.000031) × 1023 atoms are
required. This large number of particles is summarised as 1 m. In science the so-called
Avogadro-constant NA is defined as

1
NA = 6.022045 ± (0.000031) × 1023 . (3.19)
mol
Thus, the molar quantity n has its base unit:

[n] = 1 mol. (3.20)

Hence, 1 mol of carbon atoms23 have a mass of 12 g. However, if one takes the same
amount of particles, i.e. 1 mol, of a different element, the mass varies. Such as 1 mol
of oxygen molecules O2 have a mass of 32 g for instance,24 cf. Fig. 3.15. This leads
to the definition of molar mass M:
m
M= . (3.21)
n

21 Isotopes are variants of a chemical element which differ in the neutron number but having the
same number of protons in its core.
22 However, this definition is arbitrary.
23 Which equals 6.022045 ± (0.000031) × 1023 atoms.
24 This is comparable to buying fruit in a supermarket: ten apples have a greater mass than ten

grapes!
54 3 System and State

Fig. 3.15 Definition of molar mass

Table 3.2 Molar masses of relevant elements


kg
Element Chemical symbol Molar mass M in kmol
Carbon C 12
Hydrogen H 1
Oxygen O 16
Nitrogen N 14
Sulphur S 32

In case of carbon atoms and oxygen molecules it follows:

g kg
MC = 12 = 12 (3.22)
mol kmol
respectively
g kg
MO2 = 32 = 32 . (3.23)
mol kmol
The molar masses of some relevant elements are listed in Table 3.2. Based on the
principle of mass conservation the molar mass of chemical bonds can be treated as
a modular system, e.g.

kg
MSO2 = 1 · MS + 2 · MO = 64 . (3.24)
kmol
Now that it is known what the molar quantity n is, the following molar state values
can be easily defined:
 3
• Molar volume vM = Vn , molm
 J 
• Molar internal energy u M = Un , mol
 J 
• Molar enthalpy h M = Hn , mol
 
• Molar entropy sM = nS , molJ K
Chapter 4
Thermodynamic Equilibrium

The principle of thermodynamic equilibrium is essential for further understanding of


thermodynamics. Systems that are in thermodynamic equilibrium do not change their
state without external influence, i.e. they are fully at rest and the state values do not
alter in time. This requires a mechanical equilibrium, i.e. the pressure respectively
forces in the system must be balanced, a thermal equilibrium, i.e. the temperature
must be balanced, and a chemical equilibrium, i.e. every chemical reaction comes to
a standstill.

Theorem 4.1 Every fully closed thermodynamic system strives in a state of thermal,
mechanical and chemical equilibrium!

Space is an example for a fully-closed system that has not yet reached thermodynamic
equilibrium: Every change of state in the interior causes space to strive towards
thermodynamic equilibrium. This is associated with the generation of entropy, as
Sect. 14.4 is going to show.
After reaching thermodynamic equilibrium the system is characterised by, see
also Fig. 4.1:
• the pressure potential, that causes forces respectively a fluid flow, disappears, i.e.
dp = 0
• the temperature potential, that causes a heat flux, disappears, i.e. dT = 0 and
• the chemical potential, that is responsible for mass transfer respectively material
k
conversation, see Sect. 24.4, disappears as well, i.e. μi dn i = 0.
i=1

Thermodynamic equilibrium is further investigated once the entropy has been intro-
duced, see Chap. 14.

© Springer Nature Switzerland AG 2022 55


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_4
56 4 Thermodynamic Equilibrium

(a) (b)

Fig. 4.1 a Imbalanced system, b System in thermodynamic equilibrium

4.1 Mechanical Equilibrium

As known from mechanics a body/system is accelerated as long as the sum of all


forces is different to zero. A system is in rest, in case all forces are balanced, i.e.

F = 0. (4.1)

A stone in a gravitational field falls as long as it hits the ground, for example, i.e.
the contact force balances the weight force. Obviously, the stone strives for a state
of equilibrium.
Mechanical Equilibrium in Closed Systems
In the case of a closed thermodynamic system, see Fig. 4.2, the internal pressure can
also be obtained from a balance of forces.
In a state of equilibrium, there must be a balance of forces so that Eq. 4.1 is
fulfilled at every point. Otherwise, the resulting force would accelerate the system
and consequently the system would not be in a state of equilibrium. The system,

Fig. 4.2 Calculating the


pressure in equilibrium state
4.1 Mechanical Equilibrium 57

indicated by the dashed line in Fig. 4.2, is assumed to be homogeneous.1 The state
value pressure p is thus uniform within the system. However, a balance of forces
can be performed at any position, so e.g. at the interface gas/piston. The downward
force is composed of the weight force of the piston and the force acting on the piston
surface due to the ambient pressure, i.e.

Fdown = m P g + penv A . (4.2)


  
Fenv

The pressure p inside the system causes an upward force on the bottom of the pis-
ton, i.e.
Fup = p A. (4.3)

From the balance of both forces

Fdown = Fup (4.4)

it follows, that
mPg
p = penv + (4.5)
A

4.2 Thermal Equilibrium

The state value temperature T is essential for quantifying the internal state of a sys-
tem. Before explaining the temperature itself, it is advisable to first explain the mean-
ing of thermal equilibrium. The idea of thermal equilibrium is outlined in Fig. 4.3.
Just imagine two separate systems: System A is supposed to be cold, while system
B is supposed to be hot. Both systems are closed systems, i.e. no exchange of mass
is possible. Nevertheless, the two systems—still separated by a dividing wall—can
get into contact with each other. Experience shows that both systems become ther-
mally equalised after some time, i.e. A and B are warm: Systems A and B have both
changed their thermodynamic state. Obviously, energy has been transferred between
the two systems.2 After a while both systems have the same temperature. During this
transient process, heat is transferred—system A is supplied with heat while system B
releases heat. This process is obviously driven by a temperature potential. The state
that both finally reach and which they cannot leave by themselves is called thermal
equilibrium—the temperature potential has disappeared and with it the drive for the
process.

1 This is an assumption. Actually, the pressure inside the system at the bottom is larger than at the
interface gas/piston due to gravity. This effect is neglected though. However, this phenomena is
investigated in Problem 4.2.
2 This thermal energy has already been introduced as heat, see Sect. 2.2.
58 4 Thermodynamic Equilibrium

Fig. 4.3 Thermal


equilibrium

Fig. 4.4 Zeroth law of


thermodynamics

Temperature as Thermal State Value—Zeroth Law of Thermodynamics

The idea of thermal equilibrium is of such importance in thermodynamics that it


was postulated in the zeroth law of thermodynamics, which is shown in Fig. 4.4 and
explained below.

Theorem 4.2 Two systems (A, B), that are respectively in a thermal equilibrium
with a third system (C), are also in a thermal equilibrium with each other.

So if there is a thermal equilibrium between A and C respectively B and C, then


there must also be an equilibrium between A and B.
What does all this have to do with the state value temperature T ?
Temperature is a measure of the thermal state of a system. Different thermal states
can be determined by the temperature.
Theorem 4.3 Systems in thermal equilibrium have the same temperature.
In contrast, systems that are not in thermal equilibrium with each other have different
temperatures. Temperature measurement is a quantitative measurement based on
thermal equilibrium, cf. Fig. 4.5. It requires a calibrated measuring device. However,
the zeroth law of thermodynamics is the basis for any temperature measurement.
With regard to Fig. 4.4, the thermometer represents system A immersed in system
B, i.e. in ice water of a known temperature. After some time, A and B are thermally
balanced. When this thermometer subsequently comes into contact with system C,
4.2 Thermal Equilibrium 59

Fig. 4.5 Principle of temperature measurement

which is a cup of tea or coffee, a transient equalisation process takes place. For
any measurement, a state of equilibrium must first be reached3 : If systems A and
C are not balanced, the measurement is incorrect. Obviously, according to Fig. 4.5,
systems A and C do have different temperatures. A thermodynamic investigation of
this example follows once the first and second law of thermodynamics have been
introduced, see Example 14.2.

4.3 Chemical Equilibrium

If chemical reactions occur in the system, a thermodynamic equilibrium additionally


includes a chemical equilibrium. For a chemical reaction, a chemical potential γ μ
is needed, analogous to a temperature potential T for a heat flow or a pressure
potential p for a volume flow. Details are discussed in Sect. 24.4, see Example 24.6.

4.4 Local Thermodynamic Equilibrium

As already mentioned, every fully closed system moves itself in a state of equilibrium.
The question now is what happens when the system is not fully closed, see Fig. 4.6,
which shows a heat-conducting wall. The system is not fully closed, as a heat flux
enters and leaves the system, but is supposed to be in steady state. Due to

Q̇ 1 = Q̇ 2 (4.6)

incoming and leaving energy fluxes are equal—the prerequisite for steady state.
Steady state means actually that no state value within the system changes in
time. According to the previous considerations, however, the system is not in

3 Hence, this requires t → ∞, see the very right sketch of Fig. 4.5.
60 4 Thermodynamic Equilibrium

Fig. 4.6 Imbalanced


systems in steady
state—example of a heat
conducting wall

thermodynamic or thermal equilibrium because the temperature distribution in the


system is not uniform, i.e. the system is heterogeneous.4 This is due to a temperature
gradient within the system that is required to realise heat fluxes into and out of the
system.5 However, the thermal imbalance is maintained by external fluxes crossing
the system boundary. Though the entire system heat conducting wall in Fig. 4.6 is not
in thermodynamic equilibrium, it can be divided into several, differential elements.
Each of these elements is in a local thermodynamic equilibrium, see [17].

Theorem 4.4 The definition of a local thermodynamic equilibrium allows us to


apply the laws of thermodynamics also to locally varying states.

4.5 Assumptions in Technical Thermodynamics

Figure 4.7 shows the assumptions applied in technical thermodynamics:


• Due to gravity, the gas pressure varies with the height of the system. However, a
constant pressure is assumed above the height6 resulting from an equilibrium of
forces directly below the piston, i.e.:
mPg
p = penv + . (4.7)
A
This assumption is further investigated in Problem 4.2.
• In order to transfer heat Q into or out of a system a temperature gradient is
required, see Fig. 4.7. This temperature gradient is ignored in common engineer-
ing approaches: The system is assumed to be in thermal equilibrium, indicated by
a uniform temperature T in the gas.

4 The fluxes at the system’s border indicate, that the system is permanently forced from outside to
be imbalanced with the environment.
5 This is part of the lecture Heat and Mass Transfer.
6 Thus, the weight of the gas is neglected.
4.5 Assumptions in Technical Thermodynamics 61

Fig. 4.7 Assumptions in thermodynamics

Fig. 4.8 Example thermodynamic equilibrium

Example 4.1 In state (1) a cylinder/piston system filled with an ideal gas, see
Fig. 4.8, has a smaller temperature T1 than the environment (Tenv = 300 K, penv =
1 bar). The mass of the piston/spring shall be m p = 20 kg, the diameter is d = 0.1 m.
The spring with a spring constant of K Sp = 0.11 kN m
is not tensioned in state (1).
Since the temperature of the gas is smaller than ambient temperature, the system
is subject to a thermal balancing process.7 As a result, heat Q is transferred into the

7 Consider that temperature is an intensive state value that can initiate a change of state!
62 4 Thermodynamic Equilibrium

Fig. 4.9 Sketch to Problem 4.1

system as long as a temperature difference is present as a process driver. The system


heats up, the gas expands, so that the piston is eventually lifted.
Finally, the system comes to a thermodynamic equilibrium, i.e. T2 = Tenv and the
system is at rest. At this point the piston has moved by a vertical distance of x. The
pressure of the gas in state (2) is now required8 : As the system is in a mechanical
equilibrium, a balance of forces can be applied at the interface of gas/piston:

FP + FSp + Fenv = F2 (4.8)

i.e.
m P g + x K Sp + penv A = p2 A. (4.9)

Solving for p2 results in

mPg x K Sp
p2 = penv + + = 1.5619 bar (4.10)
A A
with π 2
A= d = 0.0079 m2 . (4.11)
4

Problem 4.1 A U-pipe, that is attached to a low pressure gas conduit, is filled with
silicone oil (ρs = 1203 mkg3 ) as sealing liquid. To enlarge the pressure display, water
(ρw = 998 mkg3 ) is filled above the sealing liquid in the open flank (Fig. 4.9).

(a) What are the absolute pressure and the gauge pressure in the gas conduit, if
z 1 = 147 mm, z 2 = 336 mm and penv = 0.953 bar is applied?
(b) What level difference z would occur at the same pressure in the gas conduit,
if no water was filled into the U-pipe?

8 A uniform pressure inside the cylinder is assumed, i.e. the weight of the gas is ignored.
4.5 Assumptions in Technical Thermodynamics 63

Fig. 4.10 Solution to


Problem 4.1

Solution

Since the system is in equilibrium state, all forces need to balance at any position. If
the forces are not balanced, there would be an acceleration, i.e. the system could not
be in rest. So the position marked with the area A can be picked, cf. Fig. 4.10:

Fdown = Fup (4.12)

with
Fdown = ( penv + ρw g z 2 ) A. (4.13)

On the same horizontal position the pressure within the silicone oil is constant. This
leads to: 
Fup = pg + ρs g z 1 A. (4.14)

Finally one gets

pg = penv + ρw g z 2 − ρs g z 1 = 0.96855 bar (4.15)

The gauge pressure then is

p = pg − penv = 0.0155 bar (4.16)

If there was no water, the difference z would be

pg − penv
z = = 0.1317 m (4.17)
ρs g

Problem 4.2 A basin has a height of z B = 15 m. The gauge pressure at the upper
end of the basin is measured by a U-pipe manometer and equals a water column
of z M = 1600 mm against an atmosphere of penv = 1020 mbar (ρw = 1000 mkg3 ,
g = 9.81 sm2 ).
64 4 Thermodynamic Equilibrium

(a) Determine the pressure p1 at the upper end and the pressure p2 at the bottom of
the tank. The tank is filled with nitrogen. For the dependency of the density of
nitrogen on pressure it is assumed that

kg
ρN2 = C · p with C = 1.149 (4.18)
bar m3
(b) What is the pressure p2 if the tank is filled with oil instead of nitrogen? The
pressure p1 shall be the same as before. Oil is considered an incompressible
liquid with ρoil = 870 mkg3 .

Solution

The pressure p1 can be calculated by a balance of forces at position (A)

p1 = penv + ρw g z M = 1.17696 bar (4.19)

Since the density is not constant, but a function of pressure, p2 can not be calculated
as easily as p1 . To get an overview, it is wise to start with a differential approach,
see Fig. 4.11. Performing a balance of forces at the lower edge of the differential
element leads to:
p A + ρN2 g A dz B = ( p + d p) A . (4.20)
     
Fdown Fup

Thus, it yields
d p = ρN2 g dz B . (4.21)

Fig. 4.11 Solution to Problem 4.2


4.5 Assumptions in Technical Thermodynamics 65

Replacing ρN2 = C p brings


d p = C pg dz B . (4.22)

This differential equation can be solved by separating the variables:

dp
= Cg dz B (4.23)
p

and integrating from (1) to (2)


2
2
dp
= Cg dz B . (4.24)
p
1 1

Solving the integral


p2
ln = Cg z B (4.25)
p1

and rearranging finally leads to

p2 = p1 eCg zB = 1.179 bar (4.26)

In case incompressible oil is used instead of nitrogen, the density would be constant.
It follows that
p2 = p1 + ρoil g z B = 2.457 bar (4.27)
Chapter 5
Equations of State

Chapter 3 has shown how to quantify the internal state of a system. Basically, a
distinction has been made between thermal and caloric state values. Since not all
state values can be measured directly,1 equations are required to calculate them. The
paddle wheel experiment, for example, see Sect. 2.4.1, has shown that the internal
energy describes the energetic state of a fluid. However, it is not possible to measure
the internal energy of a system, so its dependency on other state variables is of interest.
Furthermore, the question arises as to how many state variables are necessary to
unambiguously define a system. Once a system is uniquely given, it must certainly be
possible to determine the other state variables. Equations of state capture the physical
dependency between the state values. Preliminary considerations of equations of state
are presented in this chapter.

Theorem 5.1 The characterisation of the state of a system by state variables is only
unambiguous if the system is in a steady state or in a (local) thermodynamic equi-
librium. Otherwise, the system would be subject to a transient equilibrium process,
i.e. the state of the system changes in time.2

5.1 Gibbs’ Phase Rule

The basis for understanding the phase rule is the principle of thermodynamic equi-
librium, which has been explained in Chap. 4. At first, it has to be clarified how many
state variables are necessary to unambiguously determine the state of an equilibrated

1 Such as enthalpy, internal energy and entropy for instance.


2 For example, the gas in a cylinder that is not in equilibrium is spatially different and the cause of
an internal balancing process. The specification of a representative pressure is not possible due to
the spatial distribution.
© Springer Nature Switzerland AG 2022 67
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_5
68 5 Equations of State

system. According to Gibbs’ phase rule3

F =C − P +2 (5.1)

the degree of freedom F depends on the number of components C as well as on


the number of phases in thermodynamic equilibrium P. The degree of freedom F
denotes the largest number of independent intensive variables, see Sect. 3.2.5, which
can be varied simultaneously and arbitrarily without influencing each other.

Thermodynamic proof for a balanced system without chemical reactions

The proof follows according to [18]. In case several components C exit in a single
phase4 P = 1, e.g. gaseous, one can specify its composition with the molar fraction5
xi of each component. Due to
 n
xi = 1 (5.2)
i=1

only (C − 1) molar fractions6 are needed to indicate the composition—the missing


mole fraction results from Eq. 5.2. Combined with pressure p and temperature T ,
there are consequently
(C − 1) + 2 = (C + 1) (5.3)

variables that describe the system in thermodynamic equilibrium.

Let us now assume that there are several phases P. According to our considera-
tions, the total number of variables to specify such a thermodynamic system is

P(C + 1). (5.4)

In the following, a system of equations is derived to determine these variables: Note


that pressure p is associated with a mechanical equilibrium and temperature T with
a thermal equilibrium. A temperature potential describes a potential for a heat flux,
a pressure potential a potential for a volume flux. Analogously, a chemical potential
describes a potential for a chemical reaction respectively diffusion, i.e. mass transfer
of a component between the phases, e.g. evaporating water from a liquid surface.

3 Josiah Willard Gibbs (11 February 1839 in New Haven, Connecticut, 28 April 1903 in New
Haven, Connecticut).
4 E.g. air is an a gaseous state, i.e. P = 1, and it contains, simplified, C = 2 components, i.e.

nitrogen and oxygen.


5 The molar fraction is introduced in Part II and is defined as x = n i . Thus, it has the meaning
i n total
of a concentration of a component in a mixture.
6 Molar fractions, i.e. concentrations, are discussed more detailed in Part II. A concentration poten-

tial, however, initiates a balancing process, e.g. dispersion of a pollutant load in air. Thus, the
concentration can be treated as an intensive state value.
5.1 Gibbs’ Phase Rule 69

In case of a thermodynamic equilibrium there is no pressure as well as no temperature


potential. Furthermore,7 the chemical potentials μi of each component in all phases
are equal.
Theorem 5.2 In a thermodynamic balanced system, there is neither a tempera-
ture/pressure potential nor a chemical potential.
Humid air is an example, that shows the impact of a chemical potential: Let us assume
humid air, i.e. an ideal mixture of dry air and vapour, is exposed to a reservoir of liquid
water. As long as the humid air is not saturated with vapour, i.e. a chemical potential
between the liquid water’s surface and the vapour phase exists, water evaporates from
the reservoir to the air, cf. Sects. 20.1.3 and 24.4.1. A thermodynamic equilibrium is
reached as soon as the potential between liquid and gaseous phase disappears. The
mass flux of liquid/vapour then stops, the air is finally in saturated state.
In thermodynamic equilibrium,8 the equations for two phases and n components
are thus given as:
TP=1 = TP=2 (5.5)

p P=1 = p P=2 (5.6)

μ P=1,C=1 = μ P=2,C=1 (5.7)

μ P=1,C=2 = μ P=2,C=2

..
.

μ P=1,C=n = μ P=2,C=n

For m phases it is
TP=1 = TP=2 = · · · = TP=m (5.8)

p P=1 = p P=2 = · · · = p P=m

μ P=1,C=1 = μ P=2,C=1 = · · · = μ P=m,C=1

μ P=1,C=2 = μ P=2,C=2 = · · · = μ P=m,C=2

..
.

μ P=1,C=n = μ P=2,C=n = · · · = μ P=m,C=n

7 Thermodynamic equilibrium means chemical equilibrium as well.


8 Phase equilibrium!.
70 5 Equations of State

The conclusion is, that there are9

(C + 2)(P − 1) (5.9)

of these mathematical equilibrium conditions. Finally, the degree of freedom F is


defined as the total number of state variables reduced by the number of equilibrium
correlations:
F = P(C + 1) − (C + 2)(P − 1) . (5.10)
     
A B

In this equation A represents the number of required variables, whereas B quantifies


the number of equations. In case F = 0 the number of variables equals the number of
equations, i.e. only one solution exists. In case F > 0 there are more variables than
equations, so that the system of equations is underdetermined. F then represents the
number of variables that can be freely chosen. Anyhow, Eq. 5.10 simplifies to

F =C +2− P . (5.11)

In the following Sects. 5.1.1, 5.1.2 and 5.1.3, this equation is physically motivated.

5.1.1 Single-Component Systems Without Phase Change

In Part I the focus is on single component systems, i.e. C = 1, that are not subject to
a phase change, i.e. P = 1. This can be a gas, e.g. N2 , that stays gaseous throughout
the entire change of state. According to Eq. 5.1 the degree of freedom is

F =C − P +2=2 (5.12)

Thus, two independent intensive state values, namely pressure p and temperature
T , are separately and independently selectable and determine the internal state of
a system. In other words, pressure and temperature of the system can be varied
independently.

5.1.2 Single-Component Systems with Phase Change

Part II examines individual components (C = 1) that change their state of aggre-


gation, e.g. water in liquid and vapour state, i.e. P = 2. Anyhow, these fluids are
treated as real fluids. Figure 5.1, known as p, T -diagram shows the aggregate state

9 Equation 5.8 has (C + 2) rows, each row has (P − 1) equations. In case P = 2, for example, there
is only one equation per row.
5.1 Gibbs’ Phase Rule 71

Fig. 5.1 p, T -diagram for


water—Degrees of freedom

of water as a function of pressure p and temperature T .10 When water is heated at


constant pressure, its temperature initially rises while it is still in a liquid state.11 At
the boiling point the fluid starts to vaporise. Hence, water is present in liquid as well
as vaporous state, i.e. P = 2. During vaporisation the temperature stays constant,
the entire thermal energy is used for the change of aggregate state.12 According to
Eq. 5.1, the degree of freedom during evaporation is F = 1. This means, only one
intensive state value is free selectable, e.g. the pressure. It is shown in Part II, that
in this two-phase region the pressure is a function of temperature, known as vapour
pressure curve.13 However, water can even exist in three phases (liquid, solid, vapour,
i.e. P = 3) coincidently. In this case the number of freedom is

F = 1 − 3 + 2 = 0. (5.13)

Thus, none intensive state value is free selectable. This so-called triple point (TP)
exists in solely one specific point, e.g. for water at 0.01 ◦ C and 611 Pa.

10 This characteristic is explained in Chap. 18.


11 This is called sensible heat because the supply of heat energy can be measured directly with a
thermometer.
12 Thus, this thermal energy is called latent heat, since it can not be measured by an increase of

temperature.
13 You may already know that the temperature at which water starts to boil depends on the pressure.

For example, the water on Mount Everest starts boiling at a lower temperature than at sea level
because of the pressure differences.
72 5 Equations of State

5.1.3 Multi-Component Systems

Multi-component systems are dealt with in Parts II and III. For a binary mixture of
ideal gases,14 for example, the degree of freedom is

F = 2 − 1 + 2 = 3. (5.14)

Consequently, three independent intensive state values are freely selectable. For
example, in a gaseous mixture of two components, pressure p and temperature T
can be freely chosen. However, the specific volume v of the mixture can be selected
independently as well by varying the mixture ratio.15 In this case, pressure p, tem-
perature T and molar fraction x1 of one component are intensive state values, that
quantify the internal state of the system. The second molar fraction x2 is not inde-
pendent any more, since the overall balance needs to be satisfied, i.e.

x1 + x2 = 1. (5.15)

5.2 Explicit Versus Implicit Equations of State

It has been shown, that the state of a single, ideal gas (C = 1, P = 1) the degree
of freedom is F = 2, i.e. pressure and temperature can be varied independently. In
addition, the following principle of experience applies, see [5]:
• Two independent intensive/specific state values fix the internal state of a single-
component (C = 1) system without phase change (P = 1). For example, pressure
and temperature can be set independently for the gas in Fig. 5.2. The pressure p
in the gas follows
mPg
p = penv + . (5.16)
A
The temperature in thermal equilibrium equals ambient temperature in case the
system is diabatic, i.e.
T = Tenv . (5.17)

• The extensive state value quantifies the size of the system, for example by n, m,
V , H , U , S.
• If only the intensive state and not the size of a system is of interest, no extensive
state value is needed.

14 In such a case two components are present, i.e. C = 2. Both are ideal gases, i.e. only one aggregate
state occurs, so that P = 1.
15 Ultimately, this is based on the variation of the third state variable, i.e. the concentration of one

component. In a binary system, the concentration of the second component is thus automatically
present.
5.2 Explicit Versus Implicit Equations of State 73

Fig. 5.2 Independent state values

Table 5.1 Exemplary combinations of state values for single ideal gases in thermodynamic equi-
librium
State value 1 State value 2 Statement
T p truea
T v( p, T ) true
T u(T ) falseb
T h(T ) false
T s( p, T ) true
v ρ falsec
h(T ) u(T ) falsed
h(T ) s( p, T ) true
p s( p, T ) true
p v( p, T ) true
p u(T ) true
p h(T ) true
s( p, T ) v( p, T ) true
a True means state is unequivocal
b Falsemeans state is equivocal
c Due to ρ = 1
v
d Information regarding p is missing

According to this, combinations of independent state values for single ideal gases
are listed in Table 5.1. It also indicates which combinations of two state values are
independent. However, the physical dependency listed in this table between the state
values is discussed in the Chaps. 6 and 12. If additional information about the size
of the system is required, an extensive state value must also be given.
74 5 Equations of State

Gases in Part I, however, are treated as ideal so that Eq. 5.12 can be applied. Part I
additionally focuses on so-called incompressible liquids.16 TThese liquids retain
their specific volume v o matter how much pressure they are exposed to. According
to [5], the specific volume remains constant with respect to temperature variations
as well, i.e.    
∂v ∂v
= =0 (5.18)
∂T p ∂p T

Anyway, Eq. 5.12 can also be applied for incompressible fluids without phase change.
Thus, if a state is uniquely given by two independent intensive/specific state values,
the other state values must depend on these two independent state values. So there
must be a mathematical correlation such as

z = f (x, y). (5.19)

This correlation is called an equation of state, which depends on the properties of


the fluid. However, equations of state can be stated implicitly as well as explicitly:
• Implicit:
f (x, y, z) = 0 (5.20)

An example is the thermal equation of state in implicit notation, which is described


in Chap. 6
f ( p, v, T ) = 0 (5.21)

respectively
pv − RT = 0 (5.22)

• Explicit:
z = f (x, y) (5.23)

The thermal equation of state in an explicit notation is as follows

T = f ( p, v) (5.24)

respectively
pv
T = . (5.25)
R

16Being incompressible is an idealised model. However, in Part II real fluids are discussed, that do
not follow this assumption any more.
Chapter 6
Thermal Equation of State

In this chapter, the correlation between the thermal state values pressure p, specific
volume v and temperature T is investigated. As discussed in Chap. 3, the thermal
state values can be easily measured. Moreover, according to Gibbs’ phase law, cf.
Sect. 5.1, for ideal gases, two intensive state values, namely pressure and temperature,
determine the state of a thermodynamic system unequivocally. There must hence exist
a mathematical function for calculating the third thermal state value v:

v = f ( p, T ) (6.1)

6.1 Temperature Variations

To derive the thermal equation of state, i.e. Eq. 6.1, the following experiment carried
out by Gay-Lussac1 is first introduced: An ideal gas of known mass m is filled into a
cylinder closed by a freely moving piston, see Fig. 6.1. The cylinder is supposed to be
insulated against the environment. The pressure inside the cylinder in thermodynamic
equilibrium can be adjusted by a variation of the mass of the piston m P . For the first
part of the experiment the piston’s mass shall be m P,1 = const. With respect to the
balance of forces, cf. Sect. 3.2.1, it is:
m P,1 g
p1 = penv + . (6.2)
A
The experiment begins with an initial temperature TA , which can be determined with
a thermometer. The supply of thermal energy heats up the system so that a new

1 Joseph Louis Gay-Lussac ( 6 December 1778 in Saint-Léonard-de-Noblat,  9 May 1850 in

Paris).

© Springer Nature Switzerland AG 2022 75


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_6
76 6 Thermal Equation of State

Fig. 6.1 Gay-Lussac law—set-up

thermodynamic equilibrium is reached after a certain period of time, see Fig. 6.1. Its
state has changed, noticeable by a different temperature TB = TA as well as a larger
volume. Obviously, the gas inside the cylinder has expanded. The specific volume
can be calculated according to the following equation

VA
vA = (6.3)
m
VB
vB = (6.4)
m
Since the mass of the piston has not been changed while heating the system up, the
pressure in state (B) is the same as in state (A).2 The result of this first experiment is
plotted in a v, T -diagram, see Fig. 6.2. To obtain a sufficient resolution of the results,
the temperature can be varied in small steps that can be influenced by the amount of
heat supplied. In a next step, the entire experiment is repeated with a modified pressure
inside the cylinder. As mentioned before, this can easily be achieved by replacing
the piston, i.e. variation of its mass. Once again, the state values can be plotted in
the introduced v, T -diagram: If the pressure is decreased, the rise of volume while
heating the system gets larger. This led Gay-Lussac to the following conclusion:
Theorem 6.1 The volume of ideal gases is at constant pressure and at constant
mass directly proportional to its temperature. A gas expands when heated up and
contracts when cooled down.
According to Fig. 6.2 the mathematical function reads as:

v0 ( p)
v = v(T ) = · T = f ( p) · T (6.5)
T0

respectively
v = f ( p) · T (6.6)

2 Assuming, that the ambient pressure penv is constant as well. This is an essential assumption for
the environment: Its state is constant and homogeneous.
6.1 Temperature Variations 77

Fig. 6.2 Gay-Lussac law—results for thermodynamic equilibrium

Obviously, Tv00 is the gradient of the linear function. The slope of the straight line is a
function of the pressure p. The larger the pressure p is, the smaller the gradient is in
a v, T -diagram. The existence and the extrapolated value of the absolute zero point
of temperature can be derived from Gay-Lussac’s law as well: If one extends the
curves in Fig. 6.2, a common origin results. Since a negative specific volume does
not make physical sense, there must also be a lower limit for the temperature.

6.2 Pressure Variations

Boyle3 and Mariotte4 conducted their research independently. Similar to the Gay-
Lussac experiment, a cylinder is filled with a constant mass m of an ideal gas and
closed by a freely moving piston. The cylinder shall have thermal contact5 to the
environment: Once a thermodynamic equilibrium is reached, the temperature of the
gas is equal to the ambient temperature T1 = TA , see Fig. 6.3. The pressure in state
(A) depends on the mass of the piston m P,A and on ambient pressure, i.e.

m P,A g
pA = penv + . (6.7)
A

3 Robert Boyle ( 4 February 1627 in Lismore, Ireland,  10 January 1692 in London).


4 Edme Mariotte ( 1620 in Dijon,  12 May 1684 in Paris).
5 It is known from heat transfer that thermal contact can be described by k A > 0.
78 6 Thermal Equation of State

Fig. 6.3 Boyle-Mariotte


law—set-up

Fig. 6.4 Boyle-Mariotte


law—results for
thermodynamic equilibrium

The next step is to increase the mass of the entire piston to m P,B = m P,A + m.
As a consequence the gas pressure is
m P,B g
pB = penv + > pA . (6.8)
A
After a long period of time thermodynamic equilibrium is reached again. Thus, the
temperature is TB = TA = T1 . However, the volume in state (B) is smaller than in
state (A). Due to the constant mass of the gas it follows for the specific volume v:

vB < vA (6.9)

Specific volume v as well as pressure p can be visualised in a p, v-diagram, cf.


Fig. 6.4. The experiment is repeated with a variation of the piston’s mass to refine
the plot for a constant ambient temperature T1 .
Finally, the experiment is repeated for another ambient temperature T2 > T1 . The
results for thermodynamic equilibrium are plotted in the p, v-diagram as well, see
Fig. 6.4. Obviously, a second curve for T2 exists. The higher the temperature of the
gas, the greater its specific volume. Every curve belonging to a certain temperature
has the shape of a hyperbola. The conclusion is therefore as follows:
6.2 Pressure Variations 79

Theorem 6.2 The pressure of ideal gases at constant temperature and at constant
mass is inversely proportional to its volume. If the pressure is increased, the volume
decreases. If the pressure decreases, the gas expands.
For each T = const., the experimental data shows, that each p, v-pair follows a
hyperbola:
const. g (T )
p= = . (6.10)
v v
In another mathematical notation, Boyle-Mariotte’s law reads as follows:

pv = g (T ) (6.11)

6.3 Ideal Gas Law

From Gay-Lussac’s law


v = f ( p) · T (6.12)

it follows by multiplying with p

p · v = p · f ( p) · T. (6.13)

Comparing with Boyle-Mariotte law

p · v = g(T ) (6.14)

the right side of the Eq. 6.13 needs to be solely a function of temperature:

p · f ( p) · T = g(T ) . (6.15)

Thus, if the right side is a function of T , the left side of Eq. 6.15 as well needs to be
solely a function of T . This can only be the case if

p · f ( p) = const. ≡ R. (6.16)

Substitution in Eq. 6.13 leads to the thermal equation of state

pv = RT (6.17)

In Eq. 6.16 a constant R has been defined which is known as the specific gas constant.
For each gas, R is individual. Through this equation of state, all thermal state variables
are linked.
80 6 Thermal Equation of State

Equation 6.17 can be multiplied by the mass of the system. Considering Eq. 3.13
results in
pV = m RT. (6.18)

Replacing the mass m with the molar quantity n and the molar mass M:

pV = n M RT. (6.19)

Dividing by the molar quantity n:

pvM = M RT (6.20)

Avogadro6 was an Italian scientist, most noted for his contribution to molecular
theory known as Avogadro’s law:
Theorem 6.3 Equal volumes of gases under the same thermal conditions, i.e. tem-
perature and pressure, contain an equal number of molecules, i.e. the same molar
quantity.
Let us assume two ideal gases A and B at the same pressure p = pA = pB and the
same temperature T = TA = TB . According to Avogadro, for the same molar quantity
n = n A = n B , the volumes are equal as well, V = VA = VB . Applying Eq. 6.19 for
each gas leads to:
pV = n MA RA T (6.21)

pV = n MB RB T. (6.22)

Dividing Eqs. 6.21 and 6.22 shows that

MA RA = MB RB = const. ≡ RM . (6.23)

The product of molar mass M and individual gas constant R is therefore constant
and the same for all ideal gases! This constant is called universal gas constant RM .
Its value is
kJ
RM = R M = 8.3143 (6.24)
kmol K

The thermal equation of state can therefore also be given in the following notation

pvM = RM T (6.25)

6 Lorenzo Romano Amedeo Carlo Avogadro ( 9 August 1776 in Turin,  9 July 1856 in Turin).
6.3 Ideal Gas Law 81

Theorem 6.4 The thermal equation of state applies to a thermodynamic equilib-


rium.7

Problem 6.1 Calculate the molar volume vM of an ideal gas at standard conditions,8
i.e. ϑ = 0 ◦ C and p = 1.01325 bar!
Solution
To answer this question, the thermal equation of state is applied:

RM T
vM = . (6.26)
p

When solving the equation, pay attention to the units. It is always a good idea to use
SI units:
J
8314.3 kmol 273.15 K m3
vM = K
N
= 22.414 (6.27)
101325 m2 kmol

1 kmol of any ideal gas cover a volume of 22.414 m3 .

Problem 6.2 Calculate the mass of V = 3.5 m3 of nitrogen (N2 ) at a temperature


of ϑ = 25 ◦ C and a pressure of p = 3 bar!
Solution
To determine the mass, the thermal equation of state is given in rearranged notation

pV
m= . (6.28)
R N2 T

However, the gas constant of nitrogen RN2 is unknown. It can be determined by using
kg
the universal gas constant RM . Thus, the molar mass of nitrogen (MN2 = 28 kmol ) is
required, so that
J
RM 8314.3 kmol J
RN2 = = kg
K
= 296.94 (6.29)
MN2 28 kmol kg K

pV 3 × 105 Pa · 3.5 m3
m= = = 11.86 kg (6.30)
RT 296.94 kgJK · 298.15 K

7 This is because the underlying experimental investigations have been carried out for thermody-
namic equilibrium. Hence, Gay-Lussac as well as Boyle-Mariotte had to wait patiently until the
state of the system no longer changed during their experiments.
8 According to DIN 1343.
Chapter 7
Changes of State

So far, thermodynamic systems have been discussed and state values categorised to
quantify the internal state of systems. In this chapter, the focus is on changes of state,
i.e. bringing a system from an initial state to a new state. This requires the following
distinction:
• In this first case, the initial state is a state of equilibrium (1), i.e. a ther-
mally/mechanically/chemically balanced system. External impacts, e.g. work
and/or heat, acting on the system bring the system into a new equilibrium state
(2), see Fig. 7.1.
• In the second case, systems are involved that are not in a state of equilibrium. In the
previous chapters it was discussed that state values have a different influence on
systems: A heterogeneous distribution, i.e. a imperfection, of intensive state values
triggers a transient balancing process that forces the system into thermodynamic
equilibrium. In contrast, imperfections of non-intensive state values do not have
the power to trigger an equilibrium process, cf. Fig. 3.14.
Nevertheless, imperfections, cf. Fig. 4.1a, in temperature, pressure, concentration1
or even a position in a gravitational field force a system to reach a state of equilib-
rium without additional external influences, see also Chap. 14. Problem 2.1 is an
example of a transient balance process. At the beginning, the column filled with
water is in a state of equilibrium because the connecting valve to the second col-
umn is closed. However, as soon as the valve is opened, a mechanical imbalance
occurs between the two columns, which drives the system into the state of equi-
librium. Driver of this balancing process is gravity. Space can also be considered
a closed system without external influences. However, since it has not yet reached
equilibrium, transient balancing processes take place inside. The result of such a
balancing process is shown in Fig. 4.1b.

1 Or rather chemical potential.


© Springer Nature Switzerland AG 2022 83
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_7
84 7 Changes of State

Fig. 7.1 Change of state


from one equilibrium state to
another

Initially, the system in Fig. 7.1 is in equilibrium state (1), i.e. all state values are
constant in time. Thus,2 the thermal equation of state

pv = RT (7.1)

can be applied. Due to external impacts, e.g. by a supply of heat or work across the
system boundary, the system changes its state. After a while the system is in a new
equilibrium state (2) so that Eq. 7.1 can be applied again.
This chapter clarifies how changes of state run. This is essential for a further
understanding of thermodynamics, since process values, e.g. work that influences a
system, always depend on which path a change of state takes. So before solving a
thermodynamic problem, it is necessary to find out which path the system takes.3
Theorem 7.1 A change of state is the transition from one state of equilibrium into
another by external impacts. However, imbalanced systems, i.e. systems that are not
in thermodynamic equilibrium, are subject to transient balancing processes. Drivers
for these changes of state are always potentials of intensive state values, e.g. pressure,
temperature or concentration differences within the system. External impacts are not
required.

7.1 The p, v-Diagram

A p, v diagram is commonly used in thermodynamics to illustrate changes of state


and, for example, to assess the efficiency of thermodynamic cycles. As already shown,
see Chap. 5, pressure p and temperature T unambiguously define a thermodynamic

2 See Theorem 6.4.


3 Key to solving problems in thermodynamics is understanding the path a change of state takes!
7.1 The p, v-Diagram 85

state for an ideal gas. Since the specific volume v is given by

RT
v = f ( p, T ) = (7.2)
p

each pair of p, v unequivocally gives a thermodynamic state for ideal gases as well.
The specific volume v is plotted as abscissa, the pressure p as ordinate. In Sect. 9.2.2
the significance of a p, v-diagram even increases, as work can also be visualised
in such a diagram. Therefore, the p, v diagram is explained in the next sections
exemplarily for typical thermodynamic changes of state.

7.1.1 Isothermal Change of State

An isothermal change of state is characterised by

T = const. (7.3)

so that the temperature stays constant for the entire change of state. Consequently,
T1 = T2 also applies. What does it mean in terms of specific volume v and pressure
p? According to the thermal equation of state

RT const.
p = f (v) = = (7.4)
v v
an isotherm, i.e. a graph showing a constant temperature, represents a hyperbola in
a p, v-diagram, see Fig. 7.2. Let us assume that an ideal gas is filled into a cylinder
which is closed by a freely moving piston. Since no gas can leave the cylinder, it is
obviously a closed system. On its way from state (1) to state (2), which follows a
hyperbola, the gas is compressed by moving the piston downwards. This reduces the

Fig. 7.2 Isothermal change of state


86 7 Changes of State

volume filled by the ideal gas. Since its mass is constant, the specific volume also
decreases:
v2 < v1 (7.5)

Consequently, according to the p, v-diagram the pressure increases. This is further


indicated by the sketch on the left in Fig. 7.2. State (1) and state (2) are both in
thermodynamic equilibrium, see Chap. 4 and Sect. 5.1:
1. The system is thermally balanced, i.e. it is characterised by a uniform temperature
distribution inside the system in states (1) and (2).
2. The system is mechanical in rest.
This achieves both thermal and mechanical equilibrium. The pressure, however,
results from an equilibrium of forces, see Sect. 3.2.1. The greater pressure in state
(2) compared to state (1) is compensated by a greater mass of the piston. However,
it is assumed that the weight of the gas can be neglected so that the gas pressure in
the system is constant and does not vary with height, see also Sect. 4.5. As shown
in Problem 4.2, the mechanical equilibrium inside the gas actually causes a greater
pressure at the bottom of the cylinder than at the top due to gravity. This effect is
ignored from now on.

Theorem 7.2 The pressure results instantaneously from a local equilibrium of forces
at the piston edge and is assumed to be uniform in the system! The weight of the gas
is therefore neglected.

7.1.2 Isobaric Change of State

In this second case, the ideal gas undergoes an isobaric change of state so that

p = const. (7.6)

Consequently, p1 = p2 also applies. Once again, the gas in the cylinder is compressed
exemplary as illustrated in Fig. 7.3. In this example, the specific volume in state (2) is
smaller than in state (1). Both states are in equilibrium, i.e. both state (1) and state (2)
do not vary in time,4 see also Sect. 7.1.1. Since the mass of the piston is kept constant,
the pressure above it is also constant. What does this mean for the temperature? This
can be easily answered with the help of the p, v diagram and the thermal equation
of state pv
T = f (v) = . (7.7)
R

4 The gas as part of the system under investigation can also be in equilibrium, although it is not in
thermal equilibrium with the external environment. In this case, the gas is adiabatically isolated and
is not influenced by the environment. In this case, the temperature inside the cylinder is constant
and uniform in time!
7.1 The p, v-Diagram 87

Fig. 7.3 Isobaric change of state

Obviously, if p = const. and v2 < v1 , the temperature needs to decrease. If there is


thermal equilibrium with the environment, the change of state of the gas was initiated
by a cooling of the environment.

The isotherm in a p, v-diagram follows a hyperbola

RT const.
p = f (v) = = . (7.8)
v v
Figure 7.3 shows the isothermals as well. According to the Eq. 7.8, the higher the
temperature, the more the hyperbolas are obviously shifted upwards. This knowledge
of isotherms is essential for further understanding of thermodynamics.

7.1.3 Isochoric Change of State

Another change of state is possible when the piston is fixed. Under this premise, all
changes of state take place at constant volume, see Fig. 7.4. Since the mass of the
gas in the cylinder is constant, the specific volume is also constant, also referred to
as isochoric change of state:
v = const. (7.9)

Consequently, v1 = v2 also applies.


In the example shown in Fig. 7.4, an ideal gas is isochorically cooled in a closed
cylinder. As can be seen from the p, v diagram, state (2) is on a lower isotherm than
state (1), i.e.
T2 < T1 . (7.10)

The change in terms of pressure can also been visualised in a p, v-diagram. Obvi-
ously, the pressure decrease as well
88 7 Changes of State

Fig. 7.4 Isochoric change of state

p2 < p1 . (7.11)

This is documented in the sketch, as the mass of the piston in state (2) is smaller than
in state (1). States (1) and (2) shall be at equilibrium, see Sect. 7.1.1.

Problem 7.1 An air mass with a volume of 540 cm3 , a temperature of 15 ◦ C and a
pressure of 950 mbar is given. The volume of the gas is reduced to 400 cm3 . Dur-
ing this compression the gas is heated up to 45 ◦ C. The gas constant is R = 0.287 kgkJK .

(a) What mass does the air possess?


(b) What is the pressure of the gas after the compression?

Solution

(a) Applying the thermal equation of state to determine the mass of the gas:

pV = m RT (7.12)

Since state (1) is known, the mass can be calculated as follows

p1 V1 0.95 × 105 Pa · 5.4 × 10−4 m3


m= = = 0.62 g (7.13)
RT1 Nm
287 kg K
· 288.15 K

(b) Due to the closed system, the mass stays constant. This leads to

m RT2
p= . (7.14)
V2

Replacing the mass gives


7.1 The p, v-Diagram 89

p1 V1 RT2 V1 T2
p2 = = p1 = 1.416 bar (7.15)
RT1 V2 V2 T1

Problem 7.2 A gas tank with a freely movable upper bounding piston is filled with
500000 kg city gas. The gas occupies a volume of 760000 m3 . After withdrawing
300000 kg at constant temperature, the floating piston sinks accordingly. Determine
the new volume of the gas, and the specific volumes before and after withdrawal.
What are the densities?

Solution

How does the change of state run?


• Constant temperature → isothermal change of state
• Freely moveable piston → isobaric change of state
This can be specified in a p, v-diagram, see Fig. 7.5. In case the change of state is
isothermal and isobaric at the same time, it needs to be isochoric as well. This can
also be derived from the thermal equation of state:

RT
v= = const. (7.16)
p

V1 V2
v1 = = = v2 (7.17)
m1 m2
m2
V2 = V1 = 304000 m3 (7.18)
m1

V1 m3
v1 = v2 = = 1.52 (7.19)
m1 kg

respectively
1 kg
ρ1 = ρ2 = = 0.657 3 . (7.20)
v1 m

Fig. 7.5 Solution to


Problem 7.2
90 7 Changes of State

Problem 7.3 A pressure tank with a total volume of Vt = 8 m3 contains carbon


dioxide CO2 at a pressure of pt = 220 bar and a temperature of 293 K. How many
containers with a holding capacity of Vc = 10 dm3 and a pressure of pc = 150 bar
can be filled out of it, if the filling runs so slowly, that the temperature of the gas
does not change?
Remark: Filling is only possible until the pressure in the tank equals the pressure
in the containers. The remnant gas in the containers from before the filling can be
neglected.

Solution

Let us first answer the question, how much gas is taken out of the pressure tank.
Answering this question is possible, since state (1), i.e. the initial state, and state (2),
i.e. when filling stops, are well known:
• State (1)
pt Vt
m t,1 = (7.21)
RT
• State (2)
pt,2 Vt
m t,2 = (7.22)
RT
The refilling stops, when an equilibrium between tank and container has been
reached, i.e.
pt,2 = pc . (7.23)

The mass that needs to be refilled is therefore known, i.e.

Vt
m = m t,1 − m t,2 = ( pt − pc ) . (7.24)
RT
Each container can hold
pc Vc
mc = . (7.25)
RT
The total number of required containers is

m Vt pt − pc
n= = · = 373 (7.26)
mc Vc pc
7.2 Equilibrium Thermodynamics 91

7.2 Equilibrium Thermodynamics

The previous discussion has shown the importance of the equilibrium principle for
thermodynamic systems. Systems at equilibrium follow the thermal equation of state
for ideal gases cf. Theorem 6.4.
In contrast to idealised conditions, however, real changes of state can proceed
through unbalanced states between initial and final state. From a thermodynamic
point of view, these intermediate states are not defined unambiguously, as the thermal
equation of state can not be applied.5 If an intermediate state is not in equilibrium,
the system would, given enough time, undergo a transient balancing process6 until
equilibrium is reached, see Chaps. 4 and 14.

7.2.1 Quasi-Static Changes of State

In conventional thermodynamics, it is assumed that changes of state of a thermody-


namic system occur in infinitesimal partial steps, e.g. dv or d p—the change of state
can also be represented in a p, v diagram. It is assumed that the system is always
approximately in a state of equilibrium. The system is therefore a homogeneous
phase, see Sect. 3.1.1, and the pressure, for example, can be uniquely determined.
This quasistatic approach is shown on the left-hand side of Fig. 7.6.

Fig. 7.6 Quasi-static versus non-quasi-static change of state: Illustration in a p, v−diagram for gas
compression

5 The experiments performed by Boyle-Mariotte as well as by Gay-Lussac, see Chap. 6, have been
conducted under the premise of thermodynamic equilibrium.
6 The system would drive itself into thermal, mechanical as well as chemical equilibrium!
92 7 Changes of State

The major premise of this approach7 is that all sub-steps are in thermodynamic
equilibrium,8 i.e. pressure as well as temperature are balanced.9 The change of state
proceeds sufficiently slowly in infinitesimal partial steps. The sum of all these partial
steps forms the total change of state. In case the weight of the gas and the temper-
ature distribution within the boundary layer are neglected, cf. Fig. 4.7, pressure and
temperature within the system are uniform. Due to the equilibrium state, pressure
and temperature specify the state unequivocally. Thus, the thermal equation of state

pv = RT (7.27)

can be applied for the entire change of state from initial state (1) to final state (2)!
Theorem 7.3 In a quasi-static change of state the system behaves as a phase, i.e. a
homogeneous distribution of all state values occurs. No transient balance processes
take place!
The right side of Fig. 7.6 shows what may happen, in case the system is not at
equilibrium during the change of state from (1) to (2). If the piston compresses the
gas in the cylinder too quickly, for example, a pressure front is generated which is
initiated by the piston and travels through the gas. Within this front, the pressure is
greater than in the rest of the volume: consequently, the system is not in mechanical
equilibrium. As already pointed out, the system strives for a state of equilibrium, but
if the velocity of the piston is too high,10 equilibrium cannot be reached. Correspond-
ing to the pressure front, a temperature front can also be observed. Consequently, the
pressure respectively temperature distribution within the gas is uneven. The system
is heterogenous. Obviously, it is impossible to specify a pressure p as well as a tem-
perature T for the system. Hence, as indicated in Fig. 7.6, a well-defined function in
a p, v-diagram can not be found. Furthermore, since the system is not at equilibrium,
the thermal equation of state can not be applied. The system is not in unequivocal
state. However, equilibrium states (1) and (2) can be reached.If the system is at rest
when the external influence stops, the system strives for equilibrium.

Theorem 7.4 A quasi-static change of state passes several balanced sub-states. All
states between initial and final state can then be calculated with thermodynamic
equation of states for homogeneous systems, e.g. ideal gas law.

7 The pressure is assumed to be homogeneous. Thus, the gravitational effect has been ignored, see
Sect. 7.1.1!
8 Imagine you are taking photos of a car that is accelerating. If your camera has a very short exposure

time, each of the many photos looks like static, all the images are sharp. In this very short period
of time, the movement of the car is negligible. If the shutter speed is too slow, the pictures will not
look static at all, because the movement will result in blurred pictures.
9 As well as chemical potentials.
10 In the following Sect. 7.2.2, it is investigated what too high means!
7.2 Equilibrium Thermodynamics 93

7.2.2 Requirement for a Quasi-Static Change of State

The criteria for a quasi-static change of state is, that the microscopic balancing must
be much faster than the macroscopic movement: In terms of the example given in
Fig. 7.6, the balancing within the system, i.e. the striving for equilibrium, needs to
be faster than the macroscopic velocity of the piston. In such a case, the system has
homogenised itself before the new pressure front occurs. In Chap. 21 it is deduced
that pressure fluctuations spread with speed of sound, i.e.

a= κ RT . (7.28)

Depending on the gas and its temperature, velocities of some 300–1000 ms are not
unusual. Compared to typical velocities of the piston in technical applications, the
requirement for quasi-static behaviour is sufficiently fulfilled, i.e.

cpiston  a. (7.29)

Therefore, a quasi-static change of state is often approximately applicable and the


system can be treated as a phase without inhomogeneities.

Theorem 7.5 Thus, even fast changes of state can be quasi-static, although the
probability of internal dissipation increases with velocity, e.g. due to turbulence in
the system!

In contrast, the inflow of a gas in a vacuum, for example, is not quasi-static, since
local and temporal gradients in pressure, density, temperature and velocity occur
during the process. Both the microscopic and macroscopic velocities are the same,
namely the speed of sound. The approximation of a quasi-static change of state is
invalid. But as soon as the inflow is complete, the system homogenises and strives
for equilibrium. Once thermodynamic equilibrium is reached, the thermal equation
of state can of course be applied.
Theorem 7.6 Obviously, the equilibrium state of a system is independent of the path
the change of state takes.
Figure 7.7 shows this thesis. Starting from equilibrium state (1) one might follow
path (A), characterised by quasi-static behaviour from (1) to (2). Instead of taking
the direct path (A), one can also reach (2) by an isochoric plus isobaric change of state,
both of them being quasi-static. This is indicated as path (B) in Fig. 7.7. However,
the state of equilibrium (2) can also be reached by a non-static change of state, e.g.
by path (C), which cannot be represented unambiguously in a p, v diagram.
Theorem 7.7 All non-static changes of state can be replaced by a quasi-static
change of state. Based on the initial state and equations of state the final equilibrium
state can be predicted.
94 7 Changes of State

Fig. 7.7 Path-independent


equilibrium state

The deviation of the state values between initial and final state can be calculated, no
matter which path has been taken:

p = p2 − p1 (7.30)

v = v2 − v1 (7.31)

T = T2 − T1 . (7.32)

7.3 Reversible Versus Irreversible Changes of State

In addition to the newly introduced quasi-static approach, thermodynamics makes


an important distinction between reversible and irreversible changes of state. This
classification is important for the introduction of entropy as a new state variable
in Chap. 12. Reversible changes of state are the benchmark for all thermodynamic
cycles as well as for all thermodynamic machines. However, quasi-static and irre-
versible/reversible changes of state must be considered with one another as they are
closely related, as will be explained in this section.
Theorem 7.8 A reversible change of state can be reversed in any detail respectively
sub-step. After a thermodynamic system has changed from state (1) to state (2), the
process is reversible if state (1) can be reached again exactly. This applies not only
to the system itself, but also to the environment. Every reversible change of state is an
idealised change of state. In contrast to a reversible change of state, an irreversible
change of state means that the initial state, i.e. with regard to both the system and
the environment, cannot be reached again. Although the system can also reach its
7.3 Reversible Versus Irreversible Changes of State 95

Fig. 7.8 Wire pendulum

initial state again in the case of an irreversible change of state, the environment is
then different from before.
The following examples clarify the meaning of reversible/irreversible changes of
state. A distinction is made between mechanical and thermal problems.

7.3.1 Mechanical

Example 7.1 With the help of simple mechanical applications, the principle of
reversibility can be easily demonstrated. Figure 7.8, for example, shows the physical
behaviour of a so-called wire pendulum, as it is probably known from mechanics.
Initially, the mass is moved out of its rest position in a state (0), characterised by the
maximum amplitude x0 . Due to gravity, the pendulum starts swinging. Let us plot
the amplitude as a function of time, i.e. x(t), indicated in Fig. 7.8. After some time,
the pendulum returns to its rest position because its energy has been converted into
friction, e.g. at the mounting point of the wire and due to air resistance. Of course,
the initial state (0) can be restored: However, the system cannot do this itself, but
requires an external impact from the environment, e.g. mechanical work can be sup-
plied. Thus, the environment that provides the required energy has changed. Though
state (0) is restored, the entire process is irreversible.
If there were no friction, the pendulum would periodically return to the state (0),
see Fig. 7.8, and no changes in the environment could be measured.

Theorem 7.9 It is common experience that reversible changes of state are always
frictionless respectively free of dissipation.

However, friction, like in the previously discussed wire pendulum, can occur in
any thermodynamic system. Figure 7.9 shows a gas being compressed in a cylinder.
Due to the displacement of the piston the fluid inside moves. According to the no-slip
condition the gas at the piston’s surface has the same velocity as the piston, while
the fluid at the bottom of the cylinder is still in rest. Hence, a relative velocity among
96 7 Changes of State

Fig. 7.9 Dissipation in a


closed systems

the fluid particles occurs. The faster the piston moves, the more turbulence within
the gas occurs and so-called turbulent eddies are formed. Thus, the fluid particles are
rubbing against each other, i.e. internal friction arises.
This internal friction is denoted dissipation  and is discussed in detail in
Sect. 9.2.4. Anyhow, the movement of these eddies always consumes and never
releases energy. The energy consumed is dissipated, similar to the wire pendulum,
i.e. converted from mechanical energy to frictional energy.

Example 7.2 Figure 7.10 shows another example for an irreversible mechanical sys-
tem. In state (1) a ball is fixed at a height of z above ground in a gravitational field.
The fixing is loosened so that the potential energy of the ball is converted into kinetic
energy. Obviously, the ball is bouncing several times but after a while it finally comes
to rest at ground level. It is the friciton, i.e. with the surrounding air and due to the
impact at the ground, that lets the ball come to rest. Potential energy is increasingly
dissipated so that the amplitude becomes steadily smaller. It is certainly possible to
bring the ball back to the state (1), but this would require an external intervention,
i.e. mechanical energy must be supplied to the ball to bring it back to the energetic
state (1). Thus, the process is irreversible. Anyhow, the system is driven by z in
the gravitational field. Once, z to the ground has gone, there is no more driver to
change the state of the system.

Example 7.3 A transient balancing also occurs, if there is a vessel, filled with an
ideal gas at a pressure of p1 > penv , cf. Fig. 7.11. In case there is a small orifice in the
tank, a pressure balancing takes place, see Chap. 4. The volume flow rate is driven by
the pressure potential. Once, the system is at equilibrium, the volume flow disappears
and the system can not move itself back into state (1). Hence, this balancing process
is irreversible as well. Usually it is assumed that the environment is very large, so
that ambient pressure is assumed to be constant.
7.3 Reversible Versus Irreversible Changes of State 97

Fig. 7.10 Bouncing ball driven by z

Fig. 7.11 Pressure balancing driven by p

7.3.2 Thermal

Example 7.4 Changes of state are even irreversible when they strive for a thermal
equilibrium, see Fig. 7.12. Initially, the fluid has a temperature of T1 > Tenv . The
temperature potential T1 − Tenv causes a heat flux from fluid to environment:

Q̇ = k A (T1 − Tenv ) . (7.33)

This reduces the internal energy of the fluid, indicated by a decreasing temperature.
On the other side, the internal energy of the environment increase. Since in thermo-
dynamics the environment is supposed to be huge, its temperature is assumed to be
constant. As a result, the driving potential for this change of state decreases over
time. After some time, the temperature of the liquid approaches ambient tempera-
98 7 Changes of State

Fig. 7.12 Cooling down of a liquid driven by T

ture as long as the system is not adiabatic. This new equilibrium no longer has a
temperature difference to the environment. Thus, the temperature potential reaches
zero. As a result of this decreasing potential, the system can eventually no longer
move thermally! State (1) can only be reached if energy is supplied across the sys-
tem boundary, e.g. by heating the fluid up externally. Consequently, this process is
irreversible.

7.3.3 Chemical

Example 7.5 This example shows the ideal mixing of two non-reactive components
(A and B), cf. Fig. 7.13. According to Dalton, see Chap. 19, each component occupies
the entire available volume. Initially, in the left chamber the concentration of gas A
is ξA = 1, while it is ξA = 0 in the right chamber.11 This concentration imbalance
causes—as soon as the wall is removed—a transient balancing process until state
(2) is reached. At this point, the driving potential for the change of state comes to a
standstill and the system is not able to return to state (1) on its own. The process is
therefore irreversible.

Theorem 7.10 The examples shown have in common that all transient balancing
processes are irreversible: Every balancing process needs a potential as driver. This
can be pressure, temperature or concentration. During the balancing the potential
becomes smaller; once it is zero the balancing terminates and due to the missing
potential the system can not move back without external influences.

11 Mass fractions ξi are introduced in Part II though.


7.3 Reversible Versus Irreversible Changes of State 99

Fig. 7.13 Mixing of gases


driven by ξi

7.4 Conventional Thermodynamics

An overview of possible scenarios is given in Fig. 7.14. This example shows two
successive changes of state. Step 1 is an adiabatic compression followed by step 2,
an adiabatic expansion until mechanical equilibrium with the environment is reached.

• Case (A) according to Fig. 7.14 (cpiston → 0)

Reversible changes of state are always quasi-static as well. If a partial step were
not quasi-static (i.e. the system is not in equilibrium), a transient balancing process
would take place. Balancing processes are always irreversible, as can be seen from
Theorem 7.10.

Theorem 7.11 For a reversible process a quasi-static change of state is required!

Turbulence, which creates eddies and consequently leads to inhomogeneities, i.e.


irreversibilities, can be avoided if the velocity of the piston is very small. Strictly
speaking, dissipation occurs even at very low velocities. Case (A) is therefore only a
theoretical reference case. However, such reversible processes are technical reference
cases, see Chap. 15!

• Case (B) according to Fig. 7.14 (cpiston < a)

In this case, internal friction occurs because the gas is in motion, which is indicated
by vortices inside the cylinder. In this case, the fluid is swirled, i.e. there is internal
friction between the fluid particles. However, this internal friction is called dissi-
pation . This phenomenon is quite complex and indeed leads to inhomogeneities
within the fluid. In any case, Theorem 7.9 has shown that systems with friction can
never be reversible. Thus, strictly speaking, the system is no longer a phase and
100 7 Changes of State

Fig. 7.14 Overview irreversible/reversible—Example of an adiabatic compression/expansion

is not quasi-static. As a consequence, the equilibrium thermodynamics approach,


e.g. the thermal equation of state, cannot be applied. This is a dilemma. Moreover,
these inhomogeneities also trigger transient balancing processes. The process is
therefore irreversible. Actually, every change of state is based on a disturbance of
the quasi-static state within a system. To apply the means of thermodynamics, the
approximation is made that systems with internal, i.e. irreversible, friction behave
as phases and undergo quasi-static changes of state.
Consequently, in this quasi-homogeneous system all state values are actually given
by mean values—imperfections due to dissipation are then part of these mean values.
According to [1], this is only permissible if the inhomogeneities are not too large.
So the conclusion for case (B) is:
7.4 Conventional Thermodynamics 101

Theorem 7.12 Not all quasi-static changes of state are reversible!


Due to the approximation made, the thermal equation of state can be applied in case
(A) as well as in case (B), since the system is supposed to be in equilibrium12 in each
sub-step while the change of state is taking place. The appearance of dissipation,
i.e. small turbulence eddies, correlates with the macroscopic velocity of the change
of state. The larger the velocity, the higher the chance of dissipation. In contrast,
the smaller the velocity, the larger the chance of avoiding dissipation. Dissipation is,
as will be discussed further in Sect. 9.2.4, driven by the process and thus a process
value. Since case (A) and case (B) are supposed to be quasi-static, the corresponding
changes of state can be illustrated in a p, v-diagram. However, due to dissipation the
curves run differently, see Problem 7.4.

• Case (C) according to Fig. 7.14 (cpiston > a)

In case (C), the change of state proceeds so rapidly that thermodynamic equilibrium
cannot be reached in each individual sub-step. This results in a pressure front, as
shown in the example by the uneven distribution of the points that are supposed
to represent the gas. The change of state is not static because there is an uneven
distribution of state values. As already explained, such an imbalance causes a
transient balancing process. The system tries to reach equilibrium, see Chap. 4.
The previous examples have shown that every balancing process is irreversible.
Theorem 7.13 Non-static changes of state are always irreversible!
The thermal equation of state can not be applied for Case (C). Furthermore, a non-
static change of state also causes dissipation.
The previous examples (A), (B) and (C) and Theorem 7.10 have shown:
Theorem 7.14 Irreversibility is either due to friction13 or due to deviations from
thermodynamic equilibrium. These deviations force the system to run a transient
balancing process in order to achieve equilibrium. However, all balancing processes
in nature are irreversible!
The need of a process to be quasi-static in order to be reversible as well, can be
illustrated with a simplified example, see Fig. 7.15. The gas inside a cylinder must
first be expanded. Once the expansion is complete, the original state is to be restored.
The question is what is required to make the process reversible.
First, the piston is covered with a large number of weights. During the change of
state, weight after weight is moved into a reservoir on the right side of the cylinder.
This is done without lifting the weights against gravity. They are merely shifted in
parallel. Furthermore, there should be no friction when the weights are moved. When
the last weight is removed, the maximum expansion is reached. During the reverse
change of state, weight by weight is moved back onto the piston. When the last

12 And thus a homogeneous phase!


13 Later on, internal friction is called dissipation!
102 7 Changes of State

Fig. 7.15 Analogous model according to [19]

weight has been placed on the piston, the compression is finished and the system is
returned to its initial state. Due to the large number of weights, i.e. the quasi-static
approach, not only the gas is in initial state, but also the environment, i.e. in this case
the reservoir. The process is therefore reversible! However, reversibility requires a
fairly large number of weights and thus a large number of shelves. With a small
number of weights, the weights still have to be lifted from the outside to store them
on the shelves. This is not necessary with an infinite number of weights.
Another way to solve this problem is shown in the lower part of Fig. 7.15. Instead
of performing the change of state quasi-statically, it is done in one respectively two
steps. All the mass is removed from the piston and the gas expands immediately. To
compress the gas, another weight, which has been stored in the upper position, is
now placed on the piston. This causes the compression. The gas is thus back in its
initial state. However, the whole process is not reversible, the gas looks like before,
but the environment does not.

Conclusions

Though technical processes run irreversibly and non-static, the following approach
is applied, see Fig. 7.16.

Theorem 7.15 If irreversibility is not due to a missing thermodynamic equilibrium,


shown previously as case (C), but due to the release of frictional energy, i.e. case
7.4 Conventional Thermodynamics 103

Fig. 7.16 Approach


technical thermodynamics
according to [1]

(B), a slow, quasi-static change of state, similar to a reversible change of state, see
case (A), can be assumed, see [1].

Such cases are given by a continuous curve in a p, v-diagram for instance, see
Fig. 7.16. At each point along the curve a thermodynamic equilibrium is reached,
though friction/dissipation occurs. Consequently, the thermal equation of state is
valid for the entire process and the methods of equilibrium thermodynamics can be
applied. However, if friction occurs, i.e. case (B), the process runs differently than
in case (A). Thus, the illustration in a T, s- as well as in a p, v-diagram show these
differences, see Problem 7.4.
Example 7.6 In this example14 a closed system, that undergoes an isobaric, slow
change of state is investigated. However, it is intended for the advanced readers! A
closed system that is filled with an ideal gas is heated up, cf. Fig. 7.17a. The piston
closes the system to an environment and it can move up and down. Thus, during
the very slow change of state, the pressure shall be constant all the time.15 In this
example, the requirements for an isobaric change of state are investigated. This
example finally leads to Eq. 4.5. The first law in differential notation16 for a small
movement dz of the piston, see Fig. 7.17b, yields
 
δW + δ Q = dUP + m P g dz + m P cP dcP . (7.34)

When this equation is integrated, the conventional non-differential notation is:


  1  2 
W12 + Q 12 = UP + m P g z + m P cP,2 − cP,1
2
. (7.35)
2

14 To cope with this example a knowledge of first and second law of thermodynamics is required.
15 The system is going to be further discussed later in Problem 11.4!
16 Benefit of a differential notation is, that it shows what is ongoing inbetween initial and final

equilibrium state.
104 7 Changes of State

Fig. 7.17 Isobaric change of (a) (b)


state—differential notation

However, in this case, see Fig. 7.17b and considering all terms that exceed the system
boundary, Eq. 7.34 is as follows:

δW12,env − δW12,G + δ Q 12,P − δ Q Z = dUP + m P g dz + m P cP dcP (7.36)

This equation not only takes into account the initial state (1) and the final state (2),
but gives a general description for the entire change of state between the states (1)
and (2). The fundamental equation of thermodynamics, see Chap. 12, can be applied
for the piston:
dUP + p dV
dS = = δSi + δSa . (7.37)
T

For a solid body (dV = 0) it follows17 :

dUP δfric.  δ Q
= δSi + δSa = + . (7.38)
T T T
Applied to the discussed example:

dUP δfric. δ Q 12,P δ QZ


= + − . (7.39)
T T T T
Hence, after multiplying with the temperature T :

dUP = δfric. + δ Q 12,P − δ Q Z . (7.40)

Thus, it is only the friction and the heat at the piston’s boundary, that influence
the piston’s internal energy, i.e. its temperature. Substituting dUP in the first law of
thermodynamics, see Eq. 7.36, leads to

δW12,env − δW12,G = δfric. + m P g dz + m P cP dcP . (7.41)

17 The term friction fric. is explained in Sect. 9.2.5.


7.4 Conventional Thermodynamics 105

Now let us have a closer look at the work that crosses the piston’s system boundary.
Consequently, the partial energy equations are:
• The work at the upper boundary can easily be calculated from the view of the
environment18 :
δW12,env = − penv dV (7.42)

Ambient air is compressed, i.e. volume work is released to the environment. This
work is provided by the sketched system. As the environment is supposed to be
homogeneous there is no dissipation.
• The work at the lower boundary can be calculated from the point of view of the
gas. It is composed of volume work at the gas, dissipation within the gas and
mechanical work in order to move the centre of gravity of the gas, see Fig. 7.17a:

δW12,G = − pG dV + δ12,G + m G g dx + m G cG dcG (7.43)

If the piston moves by dz the centre of gravity of the gas moves by dx = dz


2
!
Hence, Eq. 7.41 reads as

− penv dV + pG dV − δ12,G − m G g dx − m G cG dcG =


(7.44)
δfric. + m P g dz + m P cP dcP .

With the velocity of the piston


dz
cP = (7.45)
dt
and the velocity of the gas
dx
cG = (7.46)
dt
it is
dx
− penv dV + pG dV − δ12,G − m G g dx − m G dcG =
dt (7.47)
dz
δfric. + m P g dz + m P dcP .
dt
With the acceleration of the piston

dcP
aP = (7.48)
dt
and the acceleration of the gas

18If the gas inside the cylinder expands, i.e. dV > 0, work is released by the system to the envi-
ronment. Thus, δW12,env , as assumed in the sketch, is negative. The related work can be calculated
easily since the ambient pressure is constant.
106 7 Changes of State

dcG
aG = (7.49)
dt
it follows:
− penv dV + pG dV − δ12,G =
(7.50)
δfric. + m P g dz + m P aP dz + m G g dx + m G aG dx.

Since
1
dx = dz (7.51)
2
it follows
dx 1 dz 1
cG = = = cP (7.52)
dt 2 dt 2
and
dcG 1 d2 z 1
aG = = = aP . (7.53)
dt 2 dt 2 2
This leads to

− penv dV + pG dV − δ12,G =
1 1 (7.54)
δfric. + m P g dz + m P aP dz + m G g dz + m G aP dz.
2 4
Dividing by dz:
δ12,G
− penv A + pG A − =
dz
  (7.55)
δfric. mG 1
+ m P g + m P aP + g + aP .
dz 2 2

Respectively, after dividing by A


 
δ12,G + δfric. mP 1 1
pG = penv + + (g + aP ) + m G g + aP . (7.56)
A dz A 2A 2

The dissipation follows19

δ12,G = ψ12,G dm = ψ12,G ρ A |dz| (7.57)

and
δfric. = ψfric. dm = ψfric. ρ A |dz|. (7.58)

19 For a technical application, mathematical functions for ψ12,G and ψfric. are needed. These func-
tions must correlate the specific dissipation with the distance moved and the velocity of a system.
A detailed example is given with Problem 11.5.
7.4 Conventional Thermodynamics 107

If it is now additionally assumed that m G  m P , i.e. kinetic and potential energy of


the gas can be neglected, then the following equation is obtained

  |dz| m P
pG = penv + ρ ψ12,G + ψfric. + (g + aP ) (7.59)
dz A

Consequently, the larger the dissipation is, the larger the pressure inside is during
the change of state. As mentioned in the problem description, the change of state
shall run very slowly. Under such circumstances, the dissipation disappears in this
mathematical limiting case.
Theorem 7.16 Very slow is a synonym for a process that does not cause any turbu-
lence within the system and is therefore free of dissipation.
The process runs isobarically for an ideal gas under the following premises:
• The change of state runs very slowly. According to Theorem 7.16, the dissipation
within the gas disappears, i.e.
ψ12,G = 0. (7.60)

• The piston moves frictionless, i.e.

ψfric. = 0. (7.61)

• The system is non-accelerated, i.e.

aP = 0. (7.62)

The mass of the gas is small, so that kinetic and potential energy of the gas are
ignored compared to the piston, i.e.

ea,12,G = 0. (7.63)

Under such circumstances the dissipation disappears and Eq. 7.59 is identical with
Eq. 4.5 for a balance of forces at rest:

mP
⇒ pG = penv + g = const. (7.64)
A

Thus, the pressure is constant for the entire change of state from (1) to (2)! Further-
more, Eq. 7.55 can be rearranged,20 so that

δ12,G + δfric.
m P aP = − penv A + pG A − − m P g. (7.65)
dz

20 With the mass of the gas still ignored.


108 7 Changes of State

Fig. 7.18 p, v- and T, s-diagram according to Fig. 7.14

Replacing the friction, as described above, leads to

  |dz|
m P a = − penv A + pG A − ρ A ψ12,G + ψfric. − mPg (7.66)
dz

This equation may be known from mechanics and has the form

mPa = F. (7.67)

In case not all forces equalise themselves, the mass is accelerated.

Problem 7.4 This advanced problem21 refers back to the classification of conven-
tional thermodynamics, see Example 7.6. How do the changes of state (A) and (B)
according to Fig. 7.14 look like in a p, v- respectively T, s-diagram?

Solution

Case (A) is rather simple. Since the entire process shall be adiabatic and reversible,
i.e. free of dissipation, the process is isentropic. Starting from state (1) during the
isentropic expansion the temperature rises until the final position has been reached
(2A). The way back is also isentropic, so it follows exactly the same path as for the
compression. Both paths can easily been drawn in the p, v- and T, s-diagram, see
Fig. 7.18.
Case (B) is somehow more difficult. From (1) to (2B) more work has to be supplied,
since dissipation occurs: Turbulent eddies, that consume energy, need to be supplied
with energy as well. Consequently at the final position (2B) the temperature is larger
than in state (2A). However, the specific volume is the same, since the piston shall

21 To cope with this problem a knowledge of first and second law of thermodynamics is required.
7.4 Conventional Thermodynamics 109

stop at the same position as in case (A):

v2A = v2B (7.68)

From (1) to (2B) entropy increases. This is due to friction and not due to heat, since
the process is assumed to be adiabatic. As indicated in the p, v-diagram the pressure
is greater than in state (2). On its way back during expansion the entropy further
increases, since friction still occurs. Mechanical equilibrium is achieved, when

p3B = p1 . (7.69)

Both, p, v- and T, s-diagram, show, that the volume in the final position is larger
than in state (1):
v3B > v1 . (7.70)

This is, since


T3B > T1 . (7.71)

Theorem 7.17 This problem has shown that internal friction, i.e. dissipation, is the
cause of the change of state being irreversible—similar to a conventional mechanical
problem as given with Example 7.1!

Anyhow, case (C) can not be illustrated in p, v- and T, s-diagrams, since no equilib-
rium is reached!
Chapter 8
Thermodynamic Processes

So far, it has been clarified what a thermodynamic system is and how its state can
be determined. In conventional thermodynamics, the principle of thermodynamic
equilibrium forms the basis for all thermodynamic calculations. Unbalanced systems
strive towards equilibrium. Furthermore, external influences can cause a system to
change from one state of equilibrium to another, i.e. a change of state takes place
with respect to time. The term thermodynamic process is defined as follows
Theorem 8.1 A thermodynamic process can consist of one or more successive
changes of state.

8.1 Equilibrium Process

Equilibrium processes have been described in detail in Chap. 7. Most of the tasks
and problems treated in this book are based on this principle. The prerequisite is that
the change of state follows a quasi-static process, i.e. in each partial step the system
is in equilibrium. Furthermore, in the state of equilibrium there are no discontinuities
within the system that would trigger a balancing process, see Fig. 7.14 for further
explanation.
An example of such a process is shown in Fig. 8.1. Obviously, dissipation 12 , i.e.
internal friction, occurs on the way from state (1) to state (2). In this book, friction
is indicated by the small vortex symbols within the system. Although it appears
that these dissipation vortices are local spots,1 they are evenly distributed to achieve
a state of equilibrium., cf. Theorem 7.15. The quasi-static change of state can be

1 If they were local spots, there would be discontinuities. In such a case, the system would not be
in equilibrium.

© Springer Nature Switzerland AG 2022 111


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_8
112 8 Thermodynamic Processes

Fig. 8.1 Equilibrium process

visualised as a smooth curve in a p, v-diagram, see Fig. 8.1. At any point on this
curve the thermal equation of state

pv = RT (8.1)

can be applied.

8.2 Transient State

A transient state occurs when the thermodynamic equilibrium of a system is dis-


turbed. Such processes are shown in Fig. 8.2. Note that a disturbance of the intensive
state values causes a transient balancing process, see also Chap. 7: In state (1) the
two tanks are separated from each other and do have different pressures inside, see
Fig. 8.2a. Nevertheless, each of the tanks is in its own state of equilibrium.
At t = 0, the barrier between the two tanks is removed and pressure equalisation
is triggered. As already mentioned, this balancing takes place at speed of sound.
The process is therefore not static and consequently cannot be sketched in a p, v
diagram. As long as the balancing is in progress, the thermal equation of state cannot
be applied. After a short time, however, a new equilibrium state (2) is established.
For the equilibrium states (1) and (2), the principles of conventional thermodynamics
can be applied.
Figure 8.2 shows additional transient balancing processes due to (b) gravity, (c)
temperature and (d) chemical potential.
Problem 8.1 Let us take a closer look at Fig. 8.2a. Initially, the pressure in both
vessels is pA and pB respectively. The volumes are VA and VB , both filled with the
same ideal gas. Due to a diabatic2 wall, the gas temperature in both vessels is equal

2 Diabatic is the opposite of adiabatic. The heat can thus pass through the wall.
8.2 Transient State 113

Fig. 8.2 Transient state

to ambient temperature T . What is the pressure p2 after the system has reached a
new equilibrium state (2)? In state (2), the system is also in thermal equilibrium with
the environment.

Solution

Since state (1) and state (2) are both at equilibrium, the thermal equation of state can
be applied
• State (1)
pA VA = m A RT (8.2)

pB VB = m B RT (8.3)

• State (2)
p2 (VA + VB ) = (m A + m B ) RT (8.4)

The total mass shares the total volume.

p2 (VA + VB ) = m A RT + m B RT (8.5)
114 8 Thermodynamic Processes

Combining Eqs. 8.2, 8.3 and 8.5 leads to

pA VA + pB VB
p2 = (8.6)
VA + VB

8.3 Thermodynamic Cycles

Figure 8.7 shows a thermodynamic cycle in a black box illustration. The cycle runs
within the system boundary. Later, it is examined in detail which changes of state
the fluid undergoes inside, cf. Chap. 17. However, the definition of a permanently
running circuit from the fluid perspective is:
Theorem 8.2 If a system finally reaches the initial state after several changes of
state, it is known as a thermodynamic cycle. Cycles play an important role in technical
applications (e.g. thermal engine, heat pump).
Consequently, a thermodynamic cycle consists of at least two successive changes of
state, see Fig. 8.3.
Mathematically, all changes of a state value Z neutralise during one cycle, since
once the initial state is reached again, all state values are as they were at the starting
point:  
dZ = 0 resp. dz = 0 (8.7)

Taking the specific volume for instance, see Fig. 8.3:


• Change of state (1) → (2)
Δv12 = v2 − v1 (8.8)

• Change of state (2) → (1)


Δv21 = v1 − v2 (8.9)

• In total:
Δv12 + Δv21 = 0 (8.10)

Fig. 8.3 Thermodynamic


cycle
8.3 Thermodynamic Cycles 115

Fig. 8.4 Stirling cycle

respectively 
⇒ dv = 0 (8.11)

The following rule applies to the reversibility of thermodynamic cycles:


Theorem 8.3 In case all changes of state are reversible, the entire cycle is reversible.
If one single change of state is irreversible, the entire cycle is irreversible!
As mentioned before, thermodynamic cycles play an important role in thermo-
dynamics, so the focus is on cycles in Chap. 17. However, one example of a ther-
modynamic cycle is given in Fig. 8.4. It shows a Stirling cycle that consists of four
changes of state, two isothermal and two isochoric steps.
In Chap. 15 it is distinguished between clockwise and counterclockwise cycles.
These cycles help to understand the meaning of the state value entropy. Clockwise
cycles, i.e. thermal engines, convert thermal energy into mechanical energy, while
counterclockwise cycles are known as cooling machines respectively as heat pumps.

8.4 Steady State Process

For steady state processes any state value Z within the system is constant in time
while it may differ locally.

8.4.1 Open Systems

Consequently, see Sect. 3.2.5, an open system in steady state is characterised by the
following balance of any extensive state3 value:

3 Extensive state variables are countable.


116 8 Thermodynamic Processes

Fig. 8.5 Steady state flow process

dZ  
=0= Ż in − Ż out + Ż Source (8.12)
dt
As indicated in Fig. 8.5 a steady state process can occur in open systems. For any
steady-state flow, the mass inflow is identical to the mass outflow.4

ṁ in = ṁ out (8.13)

If this were not the case, the mass within the system would vary in time. A system in
steady state is not only determined by constant extensive state values Z , but also by
temporally constant intensive state values, e.g. pressure p and temperature T . When
balancing an extensive state variable Z , one has to distinguish between conservation
values5 and non-conservation values, e.g. entropy. In case Z is a non-conservation
value, a source respectively sink6 term, i.e. Ż Source , can occur in the system!

8.4.2 Closed Systems

Closed systems can also be operated in a steady state. Figure 8.6 shows an example
of a heat conducting wall. According to the definitions made before, this is a closed
system, as no mass exceeds the system boundary.
Stationary in this case means that the state values are constant in time, although
they can vary locally. T (y) shows the temperature distribution within the wall, i.e.
the temperature at the top is greater than the temperature at the bottom. This profile is
constant in time and is the cause of heat entering the wall at both the top and leaving
at the bottom. In order to maintain this state, the energy supplied must be balanced
by the energy released in a steady-state condition, i.e.

Q̇ 2 = Q̇ 1 . (8.14)

4 Known as continuity equation.


5 Conservation values are mass, momentum and energy.
Source < 0.
6 Sink term means Ż
8.4 Steady State Process 117

Fig. 8.6 Steady state closed


system

Fig. 8.7 Steady state


thermodynamic cycle
(black box)

8.4.3 Cycles

Though thermodynamic cycles are discussed later, see Sect. 8.3 and Chap. 17, Fig. 8.7
shows a thermodynamic cycle in a black box notation. A thermodynamic cycle can
be permanent respectively cyclic. It is therefore treated as a stationary state. All
state values within the system boundary are constant in time. To achieve this, it is
clarified—after introducing the first law of thermodynamics—that the energy flows
must be balanced in steady state, i.e. that incoming energy equals outgoing energy:

Q̇ = Q̇ 0 + P. (8.15)

However, as can be seen from the temperatures T and T0 , since the system is operated
between hot and cold reservoir, there is a local temperature spread in the system. The
system therefore has a cold side and a hot side!
Chapter 9
Process Values Heat and Work

In the previous chapters, the internal state of a system has been described. A thermo-
dynamic system can change its state and several possible changes of state have been
discussed. For further understanding, it is essential to be familiar with the concepts
of reversibility/irreversibility as well as with the idea of a quasi-static change of state.
These approaches have been explained in the previous Chaps. 7 and 8: External influ-
ences can disturb the thermodynamic equilibrium of a system and force the system
to change its state.
In this chapter, the focus is on the so-called process values heat and work. They are
external influences that can force a system into a new state of equilibrium. First, the
thermal energy known as heat is introduced, although it has already been mentioned
several times. Then the thermodynamic work is examined. In Sect. 9.2 it is shown that
work in thermodynamic systems follows strictly mechanical principles. The concept
of work is applied to closed and open systems in this chapter.

9.1 Thermal Energy—Heat

Figure 9.1 has already been discussed in the context of the zeroth law of thermody-
namics, see Sect. 4.2. Systems A and B, which each have a different temperature,
bring themselves into thermal equilibrium when they come into contact, i.e. the tem-
peratures of A and B equalise after a while. To reach this equilibrium, energy must
be exchanged. This thermal energy is called heat. Obviously, heat occurs at the inter-
face of systems A and B: as long as the systems have different temperatures, the heat
crosses the interface of the two systems. As soon as the temperatures are balanced,
the heat at the interface disappears. In this example, the heat transferred is therefore
a function of time.

Theorem 9.1 Heat is an interaction between a system and its environment.

© Springer Nature Switzerland AG 2022 119


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_9
120 9 Process Values Heat and Work

Fig. 9.1 Thermal


equilibrium is achieved by
heat flux

Heat can therefore not be observed within a system. Within a system, the energetic
state has to be determined, e.g. by temperature or internal energy. Heat is an extensive
quantity:

[Q] = 1 J = 1 N m (9.1)

For flow processes it is appropriate to define a heat flux,1 i.e. the change in heat
over time:
2
δQ
Q̇ = →Q= Q̇ dt. (9.2)
dt
1

The unit of a heat flux is


  J
Q̇ = 1 = 1 W. (9.3)
s
Obviously, a heat flux represents a thermal power. As already mentioned, this is
often needed for open systems as well as for machines that operate permanently,
e.g. thermodynamic cycles. From a thermodynamic point of view, it makes sense to
define a specific notation as well, i.e.:
• Specific heat in relation to the mass of the system m:

Q
q= (9.4)
m
• Specific heat in relation to the mass flow rate of the fluid ṁ:


q= (9.5)

In a next step a sign convention for the heat crossing a system boundary needs to be
fixed. Figure 9.4 shows the convention followed in this book: The sign convention
is always from the system’s perspective, i.e. heat supply has a positive sign, whereas

1Imagine a gas turbine in a stationary state: It does not emit a single portion of heat, but it does
emit heat permanently.
9.1 Thermal Energy—Heat 121

Fig. 9.2 Sign convention heat—advice

heat release has a negative sign. Systems without any heat release/supply are called
adiabatic. This includes two different scenarios:
1. The system is adiabatic due to a perfect insulation towards its environment.
2. The system is adiabatic since system and environment have the same temperature,
i.e. they are in thermal equilibrium.
When solving thermodynamic problems, it is advisable to have a solution strategy
in mind: One approach may be to always indicate the heat with an arrow pointing
into the system,2 see Fig. 9.2a. For this, it is necessary to first make a sketch of the
concrete problem. When the thermodynamic balancing is performed, the result for
heat or work must be interpreted: If the heat is positive, the assumption of the arrow
direction was correct: The system is supplied with heat.3 If, on the other hand, the
sign of the heat in the calculation is negative, heat is released from the system.
In some cases, however, this advice must be given up. Let’s look at Fig. 9.2b for
instance. This sketch shows a heat exchanger, i.e. an open system of two separated
mass fluxes. Heat is transferred from one flow to the other. As indicated in this figure,
several system boundaries can be applied, e.g. systems A and B. The recommendation
given above can easily be applied for system B, as the arrow for the heat flux Q̇ points
into system B. This determines the direction of the flux and thus the recommendation
for system A must be violated, since for this system the arrow points outwards.
Anyhow, in this case, the heat flux is the coupling of the two systems A and B. For
one of the systems the recommendation can not be applied. This is characteristic of
coupled systems and is further investigated when the first law of thermodynamics is
introduced, see Chap. 11.
Figure 9.2c shows another application where the recommendation is broken. Ther-
modynamic cycles, e.g. the thermal engine shown, usually consider the absolute val-
ues of heat/work or heat flux/power. However, the direction of the external influences

2 This approach is also applicable to the work later on!


3 Be aware that the drawing convention in this book is always from the point of view of the system!
122 9 Process Values Heat and Work

for the cycles is taken into account as it is in reality. Therefore, all fluxes shown in
Fig. 9.2c shown, as already mentioned, are absolute values. The arrows shown corre-
spond to the actual flux direction: The heat flux Q̇ is supplied to the machine, whereas
the mechanical power P and the heat flux Q̇ 0 leave the thermal engine (TE). Later,
the first law of thermodynamics is applied to this thermal engine in steady state as
follows
Q̇ = P + Q̇ 0 . (9.6)

Physically, this equation can be interpreted such that energy fluxes into the system
are balanced by energy fluxes leaving the system.4
Furthermore, closed systems such as those shown in Fig. 9.2a, are usually tran-
sient, i.e. when the system undergoes a change of state, its state values change. In
this book, systems such as those shown in Fig. 9.2c are considered to be in steady
state. The heat exchanger, see Fig. 9.2b, i.e. a typical example of an open system,
can be studied in steady state or in transient state. In any case, the open system is in
steady state as soon as the start-up of the system is completed and the incoming and
outgoing fluxes are well balanced. In such a case, the internal state does not change
in time, although it may change locally.

9.2 Work

Next to thermal energy it is external work, that has the power to change the state of a
system. Just like heat, work can be supplied or released across the system boundary.
However, though work has already been introduced in Chap. 2, a further clarification
is given in this section. Subsequently, it is investigated, how work can be calculated
for systems operated with compressible fluids. Anyhow, mechanical principles are
applied.
Theorem 9.2 Work is an interaction between system and environment. It has, as
well as heat, the power to change the state of a thermodynamic system.

9.2.1 Definition of Work

Physically, work is the scalar product of a force and a distance, see also Chap. 2:

W = F · s (9.7)

According to this mechanical principle, several forms of work are conceivable, cf.
Fig. 9.3. In the first example, a closed system consisting of a cylinder and a movable

4 In case of steady state!


9.2 Work 123

Fig. 9.3 Definition of work

piston is considered. This system is supposed to be filled with a gas, i.e. the volume
is variable. Consequently, the fluid can expand or compress without changing its
mass. If the gas is ideal, a quasi-static change of state follows the thermal equation
of state
pv = RT. (9.8)

According to Fig. 9.3 the spring is compressed while the gas in the cylinder expands.
Obviously, following Eq. 9.7, work is released by the gas and stored as energy within
the spring:
1
W12 = kF x 2 . (9.9)
2
Since the work is related to the volume change of the fluid, it is called volume
work.
Another example relates to an open system, as shown in Fig. 9.3: A fluid flow
enters under high pressure. On the way there, a stirrer installed in the system starts to
rotate and transfers its rotational energy to a shaft leading out of the system. When
the fluid leaves the system at the outlet, the pressure has dropped compared to the
inlet. Again, mechanical principles, similar to the spring, can be applied to prove that
the system has released work. Hence, the work of the rotating shaft is proportional
to torque and speed, i.e.
124 9 Process Values Heat and Work

W12 = 2π Mn dt resp. P = 2π Mn (9.10)
t

Work related with the flow energy of an open system is called technical work. Specific
technical work wt,12 is composed of pressure work y12 , mechanical work5 as well as
of dissipated energy6 :

2
P
wt,12 = = v d p +ψ12 + ea,12 . (9.11)

1
  
y12

However, dissipation has an ambivalent character: On the one hand, dissipation even-
tually leads to a standstill of all movements. Thus, dissipation should be avoided in
technical applications. Dissipation even reduces the power output or increases the
power consumption of technical components, see Chap. 16. On the other hand, it is
obvious that technical processes do not run without dissipation or friction. Without
friction, a car could not start moving and entering a curve, for example, would be
impossible. Referring to the blades of the stirrer in Fig. 9.3, without Stokes’7 no-slip
condition, there is no boundary layer required to transfer fluid energy to the stirrer
and set it in motion.
However, the stirrer in Fig. 9.3 can even supply technical work to the system, in
case its shaft is driven from outside. Consequently, a pressure increase from inlet to
outlet would be measured.
In relation to Fig. 9.3, even another form of work is possible. A shaft drives a
generator, which begins to produce electrical energy. The mechanical work on the
shaft is thus converted into electrical energy. An electrical resistor is connected, the
current dissipates the electrical energy and converts it into heat.
Theorem 9.3 The mechanical energy that exceeds the system boundary and that
can be calculated as the scalar product of force and displacement is called work.
The SI-unit for work follows:

[W ] = 1 J = 1 N m. (9.12)

For flow processes it once again makes sense to define a power, i.e. the temporal
change of work:
2
δW
P= →W = P dt (9.13)
dt
1

5 Which refers to the change in the centre of gravity of a system.


6 A detailed explanation is going to follow in Sect. 11.3.4!
7 Sir George Gabriel Stokes (13 August 1819 in Skreen, County Sligo, 1 February 1903 in

Cambridge).
9.2 Work 125

Fig. 9.4 Sign convention for the process values heat and work

with its unit


J
[P] = 1 = 1 W. (9.14)
s
Usually specific work is defined in thermodynamics as follows:
• Specific work in relation to the mass of the system m:

W
w= (9.15)
m
• Specific technical work in relation to the mass flow rate of the fluid ṁ:

Pt
wt = (9.16)

Theorem 9.4 Work as well as heat can cross a system boundary. Both are process
variables and not state variables.

With regard to the sign of the work, the same principles apply as with heat, i.e.
work delivered by a system has a negative sign, while work supplied to a system has
a positive sign, see Fig. 9.4.

9.2.2 Volume Work

Volume work has already been introduced in the previous section, see Fig. 9.3. Due to
expansion inside a closed cylinder gas can release work. Reversely, work is supplied
to the system when a gas is compressed. In this section, an equation is derived to
calculate this volume work. Figure 9.5 illustrates the approach how to calculate this
volume work. For the application of definition

W = F · s (9.17)

it must first be clarified which force F is required for the volume change. As
mentioned in Chap. 4 only systems in thermodynamic equilibrium are regarded.
126 9 Process Values Heat and Work

Fig. 9.5 Deriving volume work—Compression (1) → (2)

Fig. 9.6 Specific volume work wv,12 —illustration in a p, v-diagram

Furthermore, for any change of state quasi-static behaviour is assumed. In the cylin-
der, the gas pressure p acts on every surface, thus also on the piston with a cross-
sectional area A. This leads to a force p A on the piston that points to the right. Thus,
in order to compress the gas, in a quasi-static change of state a force F must act to
the left, which equals8 the force on the inside of the piston, i.e.

F = p · A. (9.18)

On its way from state (1) to state (2) the system is supposed to be always in mechan-
ical equilibrium. During compression, however, the pressure p may depend on the
position of the piston, cf. Fig. 9.6, i.e. it may be variable:

p = p (x) . (9.19)

Therefore, the differential volume work δWv is now first calculated when the piston
moves by a differential distance dx, see Fig. 9.5. It is assumed that p (x) at the local
position x of the piston is constant when the piston moves only by an infinitesimal
distance dx, i.e. according to Eq. 9.17 the differential volume work is

8 Actually, the force F acting from the outside is infinitesimally greater than p A.
9.2 Work 127

δWv = F · dx = p (x) A dx. (9.20)

Due to the geometrical correlation

dV = A dx (9.21)

it is
δWv = p (x) dV. (9.22)

Since the occupied volume is a function of the position of the piston as well, due to

V
x= , (9.23)
A
the pressure function obeys
p = p (x) = p (V ) . (9.24)

Hence, it is
δWv = p (V ) dV. (9.25)

The sign of work is determined by the differential dV . In order to fulfil the sign
convention, see Fig. 9.4, Eq. 9.25 must finally be adjusted9 :

δWv = − p (V ) dV (9.26)

Three scenarios are therefore conceivable, all of which follow the designated sign
convention:
• Compression (dV < 0):
According to Eq. 9.26 work is positive,10 i.e. work needs to be supplied to the
system.
• Expansion (dV > 0):
According to Eq. 9.26 work is negative,11 i.e. work is released by the system.
• No change of volume12 (dV = 0):
According to Eq. 9.26 no work is exchanged.
Equation 9.26 is in differential notation. In order to determine the total volume
work an integration from state (1) to state (2) is required:

9 Cf. the following three scenarios.


10 Force is pointing into the system.
11 Force is pointing out of the system.
12 Isochoric change of state.
128 9 Process Values Heat and Work

2 2
Wv,12 = δWv = − p dV (9.27)
1 1

For the specific volume work a division by the mass of the system is carried out, i.e.

2
wv,12 = − p dv (9.28)
1

For deriving Eqs. 9.27 and 9.28 no assumption has been made regarding the reversibil-
ity. Thus, these equations can be applied to both reversible and irreversible changes
of state. The only requirement is that the system must be in thermodynamic equi-
librium, i.e. that the change of state is quasi-static. Hence, if the change of state is
irreversible, this is not due to imbalance but due to dissipation, see Sect. 7.4.

Visualisation in a p, v-Diagram

Let us take a closer look at the specific volume work in differential notation, which
is as follows
δwv = − p dv (9.29)

As discussed previously, any quasi-static change of state can be illustrated unequiv-


ocally in a p, v-diagram. Figure 9.6 shows a compression as well as an expansion of
a closed system. The differential work δwv for a small change of specific volume dv,
according to Eq. 9.29, can be visualised by the hatched area. For the entire volume
work the differential volume work δwv needs to be integrated from (1) → (2), i.e.
2 2
wv,12 = δwv = − p dv. (9.30)
1 1

This integral is the area beneath the curve describing the change of state from state
(1) to state (2) in a p, v-diagram. Thus, the specific volume work is represented by
the grey coloured area in Fig. 9.6. Areas in a p, v-diagram, that are on the left hand
side when moving from (1) → (2) indicate supplied work (wv,12 > 0), whereas areas
on the right hand side indicate released work (wv,12 < 0). Obviously, the quantity of
specific volume work depends on which path the system takes from (1) to (2). Con-
sequently, volume work is a so-called process value, since it is path-dependent. This
is in contrast to state values, that never depend on the direction a system takes. It is
explicit that the volume work applies both to systems with and without dissipation. If
dissipation occurs in the system, only the path from (1) to (2) is different than without
dissipation.
9.2 Work 129

Problem 9.1 Air with a temperature of T1 = 273 K and a pressure of p1 = 2.8 bar is
enclosed in a cylinder V1 = 2.47 m3 with a frictionless moveable and gas-tight fitted
piston. The inner diameter of the cylinder is D = 920 mm. Calculate the extensive
and the specific volume work, if the piston movement of s = 32 cm leads to an
isothermal compression. The gas constant of air is R = 287.11 kgJK .

Solution

The volume of the gas in state (2) is


π 2
V2 = V1 − D s = 2.257 m3 . (9.31)
4
The change of state is supposed to be quasi-static, so that the thermal equation of
state can be applied for the entire change of state. It reads, since the change of state
is isothermal, as:
pV = m RT = const. (9.32)

This yields
pV = p1 V1 = p2 V2 . (9.33)

This equation shows, that pressure is a function of volume. The smaller the volume
is, the larger the pressure is:

V1
⇒ p = f (V ) = p1 . (9.34)
V
The volume work follows

2 2 2
V1 1
Wv,12 = − p dV = − p1 dV = − p1 V1 dV. (9.35)
V V
1 1 1

Solving the integral leads to

V2
Wv,12 = − p1 V1 ln = 62 369.6 J > 0. (9.36)
V1

The application of the thermal equation of state finally yields:

V2
Wv,12 = −m RT ln . (9.37)
V1

Dividing by the mass, leads to the specific volume work


130 9 Process Values Heat and Work

Fig. 9.7 Effective work Weff —illustration and visualisation in a p, V -diagram, a Compression
b Expansion

V2 J
wv,12 = −RT ln = 7068.3 > 0. (9.38)
V1 kg

9.2.3 Effective Work

Volume work describes the work that a gas releases during expansion or the work
that is supplied when a gas is compressed. Volume work can therefore be positive or
negative. Anyhow, volume work is a process value from the gas point of view. For
technical applications, the effective work is of importance, which is shown in Fig. 9.7
and takes the environment into account as well. When the volume of the gas in the
cylinder increases, work is transferred from the gas to the piston, which starts moving,
see Fig. 9.7b. The moving piston can be utilised to operate a vehicle or a generator—
thus volume work can be converted into kinetic or electrical energy. However, as
soon as there is an atmosphere on the opposite side of the piston, the effective,
usable work differs from the volume work of the gas: In case of an expansion, the
energy of the piston can not be utilised completely in a technical application, since
energy is needed partially to compress the outer atmosphere. If the piston moves in
the other direction, i.e. when the gas in the cylinder is compressed, cf. Fig. 9.7a,
9.2 Work 131

the outer atmosphere supports this compression. In that case, the outer atmosphere
expands.13 As illustrated in Fig. 9.7, a force due to the internal gas pressure acts on
the piston from the left and an opposite pressure respectively force from the external
atmosphere acts on the piston from the right. The effective force shown in Fig. 9.7
is therefore:
Feff = ( p − penv ) A. (9.39)

In order to compress the gas quasi-statically, at least a force of Feff is required on the
piston. If the gas expands quasi-statically, a force of Feff is taken from the piston. In
order to determine the effective work, the effective force needs to be scalar multiplied
with the distance the piston is moving. Since the internal pressure p may vary with
the position of the piston, a differential approach is followed,14 i.e.

δWeff = −Feff dx. (9.40)

While the gas inside the cylinder is compressed, i.e. dx < 0, the outer atmosphere
supports. Vice versa, in case the gas expands, i.e. dx > 0, the outer atmosphere
impedes the expansion, i.e.

δWeff = − ( p − penv ) dV. (9.41)

By integration, the entire effective work is

2
Weff = − p dV + penv (V2 − V1 ) (9.42)
1

The effective work thus describes the compression or expansion from the perspective
of the piston, as indicated by the system boundary in Fig. 9.7. The p, V -diagrams
show the effective work in case of compression respectively expansion. The envi-
ronmental part of the work is indicated by

< 0 Compression
penv (V2 − V1 ) (9.43)
> 0 Expansion.

In the example according to Fig. 9.7a, the effective work for compression is smaller
than the volume work, that would be required if there was no environment. This is due
to the external support from the environment. In case of Fig. 9.7b, the effective work
released by expansion is smaller than the volume work, that would be released if there

13 Although the outer atmosphere can expand or can be compressed, the ambient pressure is constant
due to the size of the environment.
14 The negative sign results from the sign convention. In this differential step dx, the pressure p is

assumed to be constant.
132 9 Process Values Heat and Work

was no environment. This is due to the external compression of the environment. In


cyclic processes, e.g. an internal combustion engine, the work at respectively from
the environment is neglected, since it neutralises while compression and expansion.
In some cases, if the ambient pressure is high, the effective work can be positive
even during expansion. Mechanical work must therefore be supplied to the piston,
i.e. the piston must be pulled, although work is released from the point of view of
the gas. This occurs when

2
1
penv > p dV. (9.44)
(V2 − V1 )
1

Problem 9.2 A frictionless movable piston compresses an ideal gas with a volume
of V1 = 0.18 m3 and a pressure of p1 = 1340 hPa isothermally to V2 = 0.03 m3 .
Ambient pressure is penv = 980 hPa. Determine the effective work performed by the
piston rod force.

Solution

The effective, usable work can be calculated according to

2
Weff = − p dV + penv (V2 − V1 ) . (9.45)
1

In order to solve the integral it is required to understand how the pressure varies with
the volume. As the change of state is isothermal, see also Problem 9.1, it is

pV = m RT = const. (9.46)

This leads to
pV = p1 V1 = p2 V2 . (9.47)

This equation shows that the pressure is a function of volume. The smaller the volume,
the greater the pressure:
V1
⇒ p = f (V ) = p1 . (9.48)
V
The effective work can therefore be calculated as follows
V2
Weff = − p1 V1 ln + penv (V2 − V1 ) = 28.517 kJ. (9.49)
V1
9.2 Work 133

Fig. 9.8 Dissipation in closed (a) and open systems (b)

9.2.4 Systems with Internal Friction—Dissipation

Dissipation can occur in thermodynamic systems: A moving fluid causes turbulent


eddies that lead to internal friction. Figure 9.8 illustrates how these eddies are gen-
erated in closed and in open systems. In case a piston moves in a closed system,
as shown in Fig. 9.8a, the fluid velocity inside the system is not constant, so that,
due to relative movement of the molecules, turbulence increases. Consequently, if
no work15 is exchanged in an isochoric system16 for instance, no internal friction
results due to the missing fluid movement. In open systems, see Fig. 9.8b, the fluid
velocity also causes turbulence and thus internal friction. Experience shows that
this dissipation energy leads to an increase in temperature in the system. However,
internal friction, i.e. dissipation, needs a driver: the required energy can be supplied
• from the outside across the system boundary, e.g. a piston that is moved, or an
impeller that is immersed and rotates inside the system
• by changing the potential or kinetic energies, e.g. a bouncing ball which hits
ground,
• by pressure drop in an open system or
• in a fully-closed system by internal balancing processes, cf. Problem 11.5.
Energy of dissipation is defined by:


12 J
12 [J] respect. ψ12 = . (9.50)
m kg

Since dissipation depends on the path the change of state takes, it is a process value
similar to work or heat for instance. It therefore does not describe the state of the

15 E.g. work by an impeller, cf Sect. 9.2.7.


16 Isochoric means there is no volume work.
134 9 Process Values Heat and Work

system, as it only occurs when the system changes its state. So dissipation may or
may not occur. However, negative dissipation does not exist, since turbulent eddies
always consume energy, i.e.
 ≥ 0. (9.51)

Theorem 9.5 The wire pendulum experiment, see Sect. 7.3, has already indicated,
that processes with dissipation are always irreversible. Experience shows: If a pro-
cess is irreversible, the work that has to be supplied to achieve the same result as
with a reversible process increases:

Win,out = Wreversible,in,out +  (9.52)

Actually, this is due to the energy consumption of the turbulent eddies. The conse-
quences of Eq. 9.52 are as follows17 :
• Case 1: Compression (closed system)
Let us assume, that in reversible operation the work for compression is

Wreversible = 10 kJ. (9.53)

If the compression is irreversible, the energy of dissipation shall be

 = 2 kJ. (9.54)

The entire work for the compression is

W = Wreversible +  = 12 kJ. (9.55)

The energy supply is therefore greater than in reversible operation, since the dis-
sipated energy must also be supplied across the system boundary.
• Case 2: Expansion (closed system)
Let us assume, that in reversible operation the released work for an expansion is

Wreversible = −10 kJ. (9.56)

If the compression is irreversible, the energy of dissipation shall be

 = 2 kJ. (9.57)

The total released work is

17The cylinder is supposed to be horizontal, so that the potential energy of the gas does not have
an impact.
9.2 Work 135

Fig. 9.9 Left: compression to the same volume, right: expansion to the same pressure (both adia-
batic) (1) → (2) reversible, (1) → (2 ) irreversible

W = Wreversible +  = −8 kJ. (9.58)

The work released is therefore less than in reversible operation, since the dissipated
energy is taken from the work released.
These examples show that from a technical point of view dissipation should be
avoided in any case: Dissipation increases the work required for compression and
reduces the work released during expansion.
Process Value Dissipation—Example Compression/Expansion
Figure 9.9 shows an adiabatic compression as well as an adiabatic expansion of a
closed system in a p, V -diagram.18
• Adiabatic compression
Let us investigate what happens in case of an adiabatic and reversible, i.e. isen-
tropic,19 compression. In order to compress the gas inside the cylinder, work needs
to be supplied from (1) → (2). In this case, the first law of thermodynamics,
neglecting the changes in the outer energies,20 reads as:

W12 + Q 12 = U2 − U1 = mcv (T2 − T1 ) (9.59)



=0

with the partial energy equation

18 For this section the knowledge of the first law of thermodynamics is required.
19 See Sect. 13.5.
20 I.e. kinetic and potential energies.
136 9 Process Values Heat and Work

2
W12 = − p dV > 0. (9.60)
1

The total work supplied is volume work and can be visualised as a grey coloured
area in Fig. 9.9. In this case total work and volume work are identical. Following
the first law of thermodynamics, the temperature increases, i.e. T2 > T1 .

Now, from (1) → (2 ) dissipation shall occur, i.e. the process is irreversible but
shall reach the same degree of compression as before, i.e. V2 = V2 . The first law
of thermodynamics in this case is

W12 + Q 12 = U2 − U1 = mcv (T2 − T1 ) . (9.61)



=0

However, the partial energy equation follows

2
W12 = − p dV + 12 > 0. (9.62)
1

The entire work W12 consists of volume work, which is indicated in Fig. 9.9 by
the hatched area, and dissipation. According to Theorem 9.5, if dissipation occurs,
the work supplied is greater than if the process were dissipation-free:
W12 > W12 . (9.63)

In accordance with the first law of thermodynamics, it therefore states


T2 > T2 . (9.64)

Therefore, the temperature increase with dissipation is greater than in the case of a
process without dissipation. The additional work required for the dissipative pro-
cess is obviously converted into internal energy. Equation 9.62 can be rearranged,
so that
2
W12 − 12 = − p dV > 0. (9.65)
1

Thus, the hatched area does not only indicate the supplied volume work but also
W12 − 12 .
• Adiabatic expansion
Let us now examine what happens in the case of an adiabatic and reversible, i.e.
isentropic, expansion. While the gas inside the cylinder expands, work is released
9.2 Work 137

from (1) → (2). In this case, the first law of thermodynamics is as follows, neglect-
ing the change in outer energies:

W12 + Q 12 = U2 − U1 = mcv (T2 − T1 ) (9.66)



=0

with the partial energy equation

2
W12 = − p dV < 0. (9.67)
1

The entire released work is volume work and can be visualised as a grey coloured
area in Fig. 9.9. In this case total work and volume work are identical, since no
dissipation needs to be driven. According to the first law of thermodynamics, the
temperature decreases, i.e. T2 < T1 .
Now, from (1) → (2 ) dissipation shall occur, i.e. the process is irreversible but
shall reach equilibrium with the environment as before, i.e. p2 = p2 . The first law
of thermodynamics for this case yields
W12 + Q 12 = U2 − U1 = mcv (T2 − T1 ) . (9.68)

=0

However, the partial energy equation follows


2
W12 = − p dV + 12 < 0. (9.69)

1 >0
  
<0
21
Note that dissipation is partly driven by the volume work released. The entire
released work W12 < 0 is composed of volume work, that is indicated by the
hatched area in Fig. 9.9, and dissipation. According to the Theorem 9.5, if dissi-
pation occurs, the absolute value of the work is smaller than if the process were
dissipation-free:
|W12 | < |W12 | . (9.70)

Rearranging this equation leads to

W12 > W12 . (9.71)

In accordance with the first law of thermodynamics, it therefore states

T2 > T2 . (9.72)

21 Dissipation consumes energy.


138 9 Process Values Heat and Work

Therefore, the temperature increase with dissipation is greater than in the case
of a process without dissipation. The work required for the dissipative process is
obviously taken from the released volume work and converted into internal energy.
Equation 9.69 can be rearranged, so that

2
− W12 + 12 = p dV > 0. (9.73)
1

Thus, the hatched area does not only indicate the released volume work but also
−W12 + 12 .

Theorem 9.6 The energy of dissipation decreases the released work −Wout,rev and
increases the supplied work Win,rev .

2
W12 = W12,rev + 12 ⇒ W12 = − p dV + 12 (9.74)
1

9.2.5 Dissipation Versus Outer Friction

Dissipation
Let us summarise what we have learned so far about dissipation by taking a closer
look at a closed system as exemplified in Fig. 9.10a. The piston movement induces
turbulent eddies in the fluid. The energy of the vortices is transferred to the fluid
and increases its internal energy as soon as the vortices break up and release their
previously consumed energy. Following the no-slip condition at the wall according
to Stoke, the fluid movement stops, so there are no vortices directly at the wall.
However, the vortices do occur within the boundary layer: According to the radial
velocity distribution c(r ), a relative movement of the fluid occurs and turbulence is

Fig. 9.10 Dissipation in closed systems due to fluid turbulence (a), outer friction (b)
9.2 Work 139

generated. The greater the fluid movement driven by the moving piston, the more
turbulent the flow.
Thus, the limiting case of a very slow (→ no eddies!) change of state, see Fig. 7.14
and Theorem 7.16, reads as:

2
W = Wrev = − p dV (9.75)
1

respectively
 = 0. (9.76)

Outer Friction
In contrast to dissipation, outer friction can also occur, see Fig. 9.10b. Outer friction
refers to solids rubbing against each other, i.e. friction takes place between solid
surfaces. The first law of thermodynamics is for solid A:


1 2
W12,A + Q 12,A = UA + m A c2 − c12 A + h (z 2 − z 1 )A . (9.77)
2

The partial energy equation22 for W12,A yields




1 2
W12,A = Wmech,A + fric.,A = m A c − c1 A + h (z 2 − z 1 )A + fric.,A .
2
2 2
(9.78)
Finally, the first law of thermodynamics can be simplified as follows:

fric.,A + Q 12,A = UA (9.79)

Consequently, it is the heat and the outer friction that cause a temperature change.
The energy for the outer friction is supplied by the work W12,A , as shown by the
partial energy equation. The work supplied W12,A is therefore not only mechanical
work to move the body A, but obviously also partly frictional work. This frictional
work increases the internal energy of system A once it has dissipated. The outer
friction therefore has the same effect as heat, which may be additionally supplied.

Although the second law of thermodynamics and the associated specific entropy
s are first introduced in Sect. 12.2, the second law of thermodynamics is now applied
to this problem. Starting with the first law of thermodynamics according to Eq. 9.79
in differential and specific notation, i.e.

δψfric. + δq = du (9.80)

22 Solids are supposed to be incompressible, i.e. no volume work occurs.


140 9 Process Values Heat and Work

dividing by the temperature T leads to

δψfric. δq du
+ = . (9.81)
T T T

Applying the caloric equation of state23 reads as


du dT
=c . (9.82)
T T
The integration from state (1) to state (2) results in

2 2
du dT T2
= c = c ln . (9.83)
T T T1
1 1

Obviously, a path description (1) → (2) was not necessary to solve the integral.
Consequently, du
T
must be a state value. This state value is denoted specific entropy s:

δψfric. δq du
+ = = ds. (9.84)
T T
    T
δsi δsa

Finally, the second law of thermodynamics reads as:

ds = δsi + δsa (9.85)

9.2.6 Mechanical Work

To change the kinetic and potential energy of a system, work is required or released.
The amount of work can be calculated with the help of mechanics, see Chap. 2, i.e.
m 2
Wmech,12 = c2 − c12 + mg (z 2 − z 1 ) . (9.86)
2
Mechanical work occurs on systems when the centre of gravity is changed in relation
to the position in the gravitational field and its velocity. Lifting a crate of beer, see
Fig. 9.11, requires mechanical work:

W12 = Wmech,12 = mg (z 2 − z 1 ) . (9.87)

23 For a solid there is no distinction between cv and c p , see Sect. 12.4.2.


9.2 Work 141

Fig. 9.11 Illustration of


mechanical work

Fig. 9.12 Shaft work for a


closed system

9.2.7 Shaft Work

A rotating shaft supplies a closed system, which is initially in a state of equilibrium


(1), with work across its system boundary, see Fig. 9.12. Within the system, the fluid
is accelerated so that the kinetic energy of the fluid increases. The fluid movement
during the supply of work is complex and can be calculated with the help of fluid
dynamics and is not further investigated by thermodynamics.
However, the energy in the system is increased by the work supplied to the system
during the time the impeller is rotating. Part of the energy supplied is needed to
accelerate the fluid from a state of equilibrium (1). Another part is dissipated by
the internal friction between the liquid particles. The friction within the system is
responsible for the movement of the fluid finally coming to a halt after the rotation of
the shaft has stopped. The system reaches a new state of equilibrium (2). The entire
work supplied has finally dissipated, i.e. it has increased the internal energy of the
fluid and thus its temperature. For the closed system, see Fig. 9.12, the first law of
thermodynamics from (1) → (2) yields24

W12 + Q 12 = U2 − U1 + E a,12 . (9.88)

The change of the outer energies follows

24 The following equations are introduced and explained in Chap. 11ff.


142 9 Process Values Heat and Work

E a,12 = 0, (9.89)

since the centre of gravity has not changed from (1) → (2) and it is in rest in both
states. The partial energy equation for the work W12 reads as

W12 = Wv,12 + Wmech,12 + 12 . (9.90)

• The volume of the system is constant, so no volume work is performed, i.e.

Wv,12 = 0. (9.91)

• No mechanical work is supplied either, since the centre of gravity is in the same
rest state in (1) and (2), i.e.
Wmech,12 = 0. (9.92)

Consequently, the entire work W12 is dissipated:

W12 = 12 . (9.93)

Thus, it follows
Wshaft = 12 (9.94)

This equation shows, that shaft work in a closed system with a motionless fluid
can only be positive. It has never been observed that such a system drives a shaft,
i.e. W12 < 0. Thus, according to the second law of thermodynamics the process is
irreversible
Si,12 > 0. (9.95)

However, the increased internal energy, measured by a temperature rise, can only be
converted partly back into mechanical work, e.g. by a thermal engine, since just the
temperature potential towards environment can be utilised, see Sect. 16.1. Dissipation
is always related with a loss of exergy, see Sect. 16.4:

E x,v,12 = Si,12 Tenv > 0. (9.96)

9.2.8 Shifting Work

In this section, open systems, as shown in Fig. 9.13 are examined. An open system
is characterised by mass passing the system boundary. Let the inlet be designated
(1) and the outlet (2). Within the system, the pressure can be distributed so that the
pressure at the inlet is p1 and at the outlet p2 . The question now is how much work is
required to push a fluid particle into the system respectively how much energy leaves
the system with a pushed out fluid particle. A piece of mass m 1 has to be moved
9.2 Work 143

Fig. 9.13 Shifting work in


open systems

across the system boundary at the inlet (1). In order to overcome the local pressure
p1 a force F1 is required, i.e.
F1 = p1 A1 . (9.97)

The piece of mass needs to be shifted by a distance x1 to be fully moved into the
system. Consequently, the work required for this step is

Win = F1 x1 . (9.98)

Substituting the force leads to:

Win = p1 A1 x1 = p1 V1 . (9.99)

Analogue, the work at the outlet (2) can be calculated. At that point the mass m 2 ,
and thus shifting work, leaves the system:

Wout = − p2 A2 x2 = − p2 V2 . (9.100)

Summarising, the work of shifting can be described as follows

Wshift = ± pV (9.101)

respectively in specific notation:

wshift = ± pv (9.102)
144 9 Process Values Heat and Work

Fig. 9.14 Technical work—no dissipation, outer energies ignored

9.2.9 Technical Work Respectively Pressure Work

The term technical work is related to open systems, that are characterised by mass
passing a system boundary. This name actually comes from fluid mechanics and
represents the flow energy. In Sect. 21.1 it is shown that the technical work indeed
follows Bernoulli’s equation for incompressible fluids. Neglecting both dissipation
and external energies, the specific technical work wt,12 simplifies to the so-called
specific pressure work y12 , as shown in Fig. 9.14.
It is now examined how much work is required to increase the pressure of a flow,
see Fig. 9.14a, or how much work a system releases when the pressure decreases, see
Fig. 9.14b. As already mentioned, dissipation and the change of external energies
are to be ignored at this point. Under these premises, the specific technical work is

2
wt,12 = y12 = v d p. (9.103)
1

This equation is derived in Sect. 11.3.4. Anyhow, three cases can occur for the
premises made:
• The pressure rises, i.e. d p > 0, so that the technical work is positive. Work is
supplied to the system.
• The pressure decreases, i.e. d p < 0, so that the technical work is negative. The
system releases work.
• The pressure stays constant, i.e. d p = 0, so that no technical work is transferred.
Technical power is obtained by multiplication with the mass flux:

2
Pt = ṁwt,12 = ṁ y12 = ṁ v · d p. (9.104)
1

Motivation of the Pressure Work


9.2 Work 145

Fig. 9.15 Equivalent model for technical work @ pressure decrease, related to Fig. 9.14b

As already mentioned, a detailed physical proof is given in Sect. 11.3.4. However,


the idea of pressure work can also be motivated by the principles presented so far.
The approach for such an explanation is shown in Fig. 9.15.
An open system in steady state according to Fig. 9.14b is described by an equiva-
lent model realised by a cylinder/piston system: Such a system has two valves at its
bottom, one for the inlet and one for the outlet. However, the changes in the external
energies of the gas are ignored in the following considerations. Furthermore, the
entire change of state shall run reversibly and thus very slowly. In a first step the
fluid enters the cylinder, so that its pressure is constant25 : Therefore, the piston must
move upwards during the filling process and be in mechanical equilibrium with the
gas during the entire filling process. The outlet valve is closed. From the gas point
of view, see system boundary in Fig. 9.15, work is released to the piston, since the
potential energy of the piston rises. As the pressure is constant, the work exchanged
at the interface gas/piston for the filling process (E) → (1) obeys

WE1 = − p1 A z 1 = − p1 V1 . (9.105)

=Fgas

Since the aim is to lower the pressure, in a second step from (1) → (2) both the inlet
and the exhaust valves are closed while the mass of the piston is gradually reduced.
This increases the volume, depending on the process control. This step is quasi-static
with a small velocity so that the principles of equilibrium thermodynamics can be
applied and the process is also intended to be free of dissipation. The sketched p, V
diagram illustrates the work that is released from the system’s point of view. As
the piston moves upwards,26 work is released at the gas/piston interface, which is

25 According to Example 7.6 the requirements for an isobaric change of state are fulfilled.
26 While doing so, the piston’s potential energy rises.
146 9 Process Values Heat and Work

therefore negative, i.e.


2
W12 = − p dV. (9.106)
1

In the last step from (2) → (A), the exhaust valve is opened while the inlet valve is
still closed and the piston moves slowly downwards so that the released fluid keeps
its pressure constant. Once again, the shifting work to release the gas can be easily
calculated and has a positive sign as the piston moves downwards,27 i.e.

W2A = p2 A z 2 = p2 V2 . (9.107)

=Fgas

The total technical work at the gas/piston interface is therefore the sum of all sub-steps

Wt = WE1 + W12 + W2A . (9.108)

Hence, it is
2
Wt = p2 V2 − p1 V1 − p dV. (9.109)
1

Applying the product rule:

d( pV ) = V d p + p dV (9.110)

brings
2 2 2
d( pV ) = V dp + p dV. (9.111)
1 1 1

Rearranging leads to

2 2
p2 V2 − p1 V1 = V dp + p dV. (9.112)
1 1

Thus, Eq. 9.109 simplifies to


2
Wt = V d p. (9.113)
1

27 The piston’s potential energy decreases. Its energy is transferred to the gas.
9.2 Work 147

The calculation of the time derivative, since the real process, see Fig. 9.14b, runs
continuously, results as follows

2
Pt = V̇ d p. (9.114)
1

Dividing by the mass flow rate ṁ leads to

2
wt = v d p. (9.115)
1

The integral obeys28

A 1 2 A
v dp = v dp + v dp + v dp . (9.116)
E E 1 2
     
=0 =0

Hence, it finally is
A
wt,EA = v dp (9.117)
E

28 (E)→(1) as well as (2)→(A) are isobaric.


Chapter 10
State Value Versus Process Value

It has been shown in the previous chapters that it is important to distinguish between
state and process values. State values indicate the state of a thermodynamic system.
Furthermore, they can be depicted, for example, in a p, v diagram. The state of
an ideal gas is defined by two independent state values, so that each p, v-pair in a
p, v-diagram uniquely defines the state of a system, see Fig. 10.1. Consequently, the
change of a certain state value Δz can be calculated by initial and final state. State
values are therefore always independent of the path the system takes:

2
Δz = dz = z 2 − z 1 (10.1)
1

In this notation d denotes the differential of a state value. No matter, which path, i.e.
(a), (b) or (c) is followed, cf. Fig. 10.1, changes of the state values Δp and Δv are
unaffected. According to Eq. 10.1 it is

Δp = p2 − p1 (10.2)

and
Δv = v2 − v1 . (10.3)

Hence, mathematical combinations of state values lead to new state values, see [20].
The specific enthalpy h, which is formally introduced in Sec. 11.3.1, is defined, for
example, as
h = u + pv. (10.4)

Obviously, the specific enthalpy h is a mathematical combination of the state values


specific internal energy u, pressure p and specific volume v. No information about

© Springer Nature Switzerland AG 2022 149


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_10
150 10 State Value Versus Process Value

the path is required for the calculation of the state-specific enthalpy h. If the state
is unambiguously determined via u, p and v, the specific enthalpy h can be easily
calculated. Once the energy balance of an open system is discussed, the state value
specific enthalpy h is physically motivated.1 However, the specific enthalpy is a state
value and does not depend on the category of a system, i.e. it is also valid in closed
systems.2 However, not every combination of state values leads to relevant state
values with vivid physical interpretation. By multiplying the specific enthalpy by the
mass, the extensive enthalpy H is finally obtained:

H = mh = U + pV. (10.5)

In contrast to state values, process values only occur when the state of a system
changes. Furthermore, they are even path-dependent, as illustrated in Fig. 10.1.
According to Chap. 9, the volume work is represented, e.g. in a p, v-diagram, as the
area beneath the path of a change of state. Obviously, the amount of work depends on
the process, i.e. on the mathematical function p = f (v), which actually describes the
trajectory of a change of state. Specific heat q and specific work w are both process
values, as is specific dissipation ψ. The differential of a process value is indicated by
a δ to separate them clearly from state values. Consequently, they follow:

2
δw = w12 . (10.6)
1

This emphasises that work is not present in state (1) or (2), but only occurs when a
system transitions from (1) to (2):

2
δw = w2 − w1 . (10.7)
1

Theorem 10.1 State values always provide total differentials.

10.1 Total Differential

In order to prove Theorem 10.1, the definition of a total differential is first given.
As shown in the previous chapters, the state of an ideal gas is uniquely determined

1 In open systems, the specific enthalpy h represents the sum of specific internal energy u and specific

displacement work required to bring a piece of fluid into the system against the local pressure.
2 Unfortunately, the specific enthalpy in closed systems does not have such a vivid meaning as in

open systems.
10.1 Total Differential 151

Fig. 10.1 Process versus


state value

with two state values, e.g. x and y. Consequently, every other state value z must be
a function of these two state values:

z = f (x, y) , (10.8)

respectively
F (x, y, z) = 0. (10.9)

This function can be represented in a diagram, see Fig. 10.2. The question now is
how to calculate the total change in z when the function depends on two independent
variables x and y. To answer this, first the variable y is fixed so that the partial change
of z from (1) to (1 ) follows:
 
∂z
dz y = dx. (10.10)
∂x y

In a second step, from (1 ) to (2), the variable x is kept constant so that
 
∂z
dz x = dy. (10.11)
∂y x

The entire change of z is then given by the total differential


   
∂z ∂z
dz = dx + dy (10.12)
∂x y ∂y x
152 10 State Value Versus Process Value

Fig. 10.2 Total differential

10.2 Schwarz’s Theorem

Theorem 10.2 The requirement for a total differential is, that point (2) according
to Fig. 10.2 can be reached path-independently.

In order to verify Theorem 10.2, Fig. 10.3 shows a top view. State (2) can be reached
via path 1 and alternatively via path 2. The following abbreviations are introduced:
 
∂z
ξ (x, y) = (10.13)
∂x y

and  
∂z
η (x, y) = . (10.14)
∂y x

Starting from state (1), both ways to reach state (2) are now examined. From the
combination of Theorems 10.1 and 10.2 it follows that the total differential for each
state value must be path independent:
• Path 1
This path follows (1) → (1 ) → (2)

dz = η (x, y) dy + ξ (x, y + dy) dx (10.15)

• Path 2
This path follows (1) → (1 ) → (2)

dz = ξ (x, y) dx + η (x + dx, y) dy (10.16)


10.2 Schwarz’s Theorem 153

Fig. 10.3 Schwarz’s


theorem

For a state value it must be irrelevant whether path 1 or path 2 is taken,3 so that
Eqs. 10.15 and 10.16 are equal:

η (x, y) dy + ξ (x, y + dy) dx = ξ (x, y) dx + η (x + dx, y) dy. (10.17)

Rearranging leads to:

ξ (x, y + dy) − ξ (x, y) η (x + dx, y) − η (x, y)


= . (10.18)
dy dx

The left side shows the derivative of ξ to y, the right side the derivative of η to x:
   
∂ξ ∂η
= . (10.19)
∂y x ∂x y

The substitution of the previously made abbreviations leads to the so-called Schwarz’s
Theorem 10.3:
∂2z ∂2 z
= . (10.20)
∂ y ∂x ∂x ∂ y

Theorem 10.3 In case Eq. 10.20 is fulfilled, a total differential exists.

Example 10.1 Ideal gas law

RT
p = p (T, v) = (10.21)
v

3 State values are never path-dependent!


154 10 State Value Versus Process Value

• Path 1
∂p R ∂2 p R
= → =− 2 (10.22)
∂T v ∂T ∂v v
• Path 2
∂p RT ∂2 p R
=− 2 → =− 2 (10.23)
∂v v ∂v ∂T v
This leads to
∂2 p ∂2 p
= (10.24)
∂T ∂v ∂v ∂T

Conclusion: The ideal gas law is an equation of state.

Example 10.2 Specific volume work is a function of pressure and specific volume,
i.e.
w = w ( p, v) (10.25)

respectively
δw = δw ( p, v) = − p dv. (10.26)

Thus, the total differential yields


   
∂w ∂w
δw = dv + d p. (10.27)
∂v p ∂p v

The comparison of the coefficients leads to


 
∂w ∂2w
= −p → = −1 (10.28)
∂v p ∂v ∂ p

and  
∂w ∂2w
=0→ = 0. (10.29)
∂p v ∂ p ∂v

Consequently, it is
∂2w ∂2w
= (10.30)
∂v ∂ p ∂ p ∂v

Conclusion: Volume work is path-dependent and thus no equation of state. A total


differential does not exist.

Problem 10.1 Derive the so-called isobaric volumetric thermal expansion coeffi-
cient
10.2 Schwarz’s Theorem 155
 
1 ∂V
β= (10.31)
V ∂T p,m

for an ideal gas!

Solution

According to the thermal equation of state, the volume of an ideal gas is obtained as
follows
m RT
V = . (10.32)
p

For keeping pressure4 and mass constant it is


 
∂V mR
= . (10.33)
∂T p,m p

This leads to
1 mR
β= . (10.34)
V p

Finally, by applying the thermal equation of state, one gets:

1
β= (10.35)
T

4 Since the isobaric thermal expansion coefficient is asked for.


Chapter 11
First Law of Thermodynamics

In this chapter, the causality between process values and the change of state values
of a thermodynamic system is derived. The first section introduces the equivalence
principle between work and heat—an essential prerequisite for the formulation of the
law of conservation of energy, known as the first law of thermodynamics. Finally,
the first law of thermodynamics is applied to closed and open systems and ther-
modynamic cycles. Anyhow, thermodynamic cycles play an important role in the
following chapters.

11.1 Principle of Equivalence Between Work and Heat

In this section, the impact of mechanical, electrical and thermal energy on a system
is analysed, see Fig. 11.1. Assuming that the initial state (1) is in equilibrium in all
three cases, the same amount of energy is supplied to the system, though in different
forms. To clarify the initial state of the system, its internal energy

U1 = mu 1 (11.1)

is first determined. If the system is filled with an ideal gas, the specific internal energy
u 1 is only a function of the temperature, see Chap. 12. All three cases therefore have
an identical temperature ϑ1 . As a result of the energy supply, the internal state of
the system strives for a new equilibrium state, i.e. the temperature ϑ2 in state (2)
is increased compared to the initial state. It obviously makes no difference whether
thermal energy is supplied by a heating plate, mechanical energy by an impeller or
electrical energy. In each case the temperature rises. Furthermore, the temperature
ϑ2 is identical in all cases when the same amount of energy is supplied.

© Springer Nature Switzerland AG 2022 157


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_11
158 11 First Law of Thermodynamics

Fig. 11.1 Equivalence of heat and work

Theorem 11.1 Heat and work can have the same effect, e.g. temperature increase,1
on a system. Heat and work are equivalent. The same change of state can be achieved.

This conclusion is important for deriving the first law of thermodynamics, i.e.
the law of conservation of energy. Consequently, thermal, mechanical and electrical
energy can be superimposed. Their interaction causes a change of state of a system.
Obviously, according to Fig. 11.2, heat can even be converted into work. If the
system is in an environment with a higher temperature than the system itself, e.g. on
a hotplate, the gas in the system heats up and expands. Due to this expansion, the
gas transfers volume work to the piston. However, the work can also be converted
into heat, e.g. by first converting it into electrical energy. The current can then cause
a flow of electricity that dissipates at an electrical resistance and eventually releases
heat.

Theorem 11.2 Heat and work are equivalent. Heat can be generated by work and
converted into work.

1 More precisely, it should be the same change in internal energy that can be observed. When a real
fluid changes its state of aggregation, for example, the temperature can even be constant although
its internal energy changes, see Part II.
11.2 Closed Systems 159

Fig. 11.2 Converting thermal to mechanical energy

11.2 Closed Systems

In this section the first law of thermodynamics for closed systems, see Sect. 3.1.2,
is derived. However, a distinction is made between systems in rest and systems in
motion.

11.2.1 Systems at Rest

Since both thermal and mechanical energy are equivalent, cf. Theorem 11.1, the
influence of the two forms of energy is examined one after the other. First, the focus
is on the influence of thermal energy on a system at rest, see Fig. 11.3.
Due to the supply of heat, the system contains a greater amount of energy in state
(2) than in state (1). Note that the bubbles in Fig. 11.3 are meant to represent energy.
Since energy cannot be destroyed or generated out of nothing, the number of bubbles
must be constant: According to the law of conservation of energy, the quantity of
supplied heat Q 12 must therefore be equal to the change of internal energy in the
system, i.e.
Q 12 = U = U2 − U1 . (11.2)

Fig. 11.3 Closed system at rest—supply of heat


160 11 First Law of Thermodynamics

Fig. 11.4 Closed system at rest—supply of work

Fig. 11.5 Closed system at rest

In Chap. 12 the rise of internal energy is correlated with an increase of temperature


for ideal gases respectively incompressible liquids.2
In a next step, mechanical energy is now supplied to the system, see Fig. 11.4.
Here, too, the principle of conservation of energy applies, namely that the supply of
mechanical work influences the internal energy. Therefore the balance obeys

W12 = U = U2 − U1 . (11.3)

The principle of conservation of energy for a closed system at rest can obviously
be summarised according to Fig. 11.5. Heat and work can be superimposed because
they are equivalent in terms of their influence on the system, see Theorem 11.2.
The first law of thermodynamics for a closed system at rest yields

Q 12 + W12 = U2 − U1 (11.4)

The only change is in the internal energy, as the system is assumed to be at rest, i.e.
the kinetic and potential energy in relation to the centre of gravity remain unchanged.
According to Eq. 11.5, the internal energy in the initial state (1) and in the final
state (2) are relevant. Therefore, intermediate information between (1) → (2) is not
required. Consequently, the formulation of the first law of thermodynamics can be
applied independently of its reversibility as well as for quasi-static/non-static changes
of state. However, path-dependent information is part of the process values Q 12 and
W12 . The first law of thermodynamics should always be formulated in extensive

2As already mentioned, real fluids do not necessarily heat up because the so-called latent heat
causes a phase change.
11.2 Closed Systems 161

Fig. 11.6 Converting energy in a closed system

notation,3 as shown in Eq. 11.5. However, since the mass of a closed system is
constant, division by the mass m leads to the following specific notation of the first
law of thermodynamics:
q12 + w12 = u 2 − u 1 (11.5)

From time to time it is advantageous to formulate the first law of thermodynamics


in differential notation, especially when basic thermodynamic principles are derived.
The differential notation obeys:

δq + δw = du (11.6)

Note that the differentials of process values are denoted by δ, while the differentials
of state values are denoted by d, see Chap. 10. Anyhow, the integration of Eq. 11.6
leads to Eq. 11.5.
Example 11.1 The following example examines the functioning of a closed system
for generating net work. The proposed engine consists of a horizontal cylinder filled
with an ideal gas and closed by a frictionless moving piston, see Fig. 11.6.
In case (a), the gas is compressed from state (1) to state (2) as indicated in the
p, V -diagram. When compressed, the volume of the gas decreases and its pressure
and temperature increase. Furthermore, as derived in Chap. 9, the required work W12
is the grey coloured area under the curve (1) → (2) as long as the change of state is

3 It is the energy that remains constant, not the specific energy.


162 11 First Law of Thermodynamics

reversible and the changes in the external energies are ignored.4 From the system’s
point of view, this work is positive because work is required for compression. If the
piston is now released—because the pressure p2 is greater than the initial ambient
pressure p1 —the gas expands. If it is assumed that it takes the same path as the
previous compression, the work W21 released during the expansion from (2) → (1)
is equal to the work that had to be expended for the compression, i.e.

W12 = |W21 |. (11.7)

Obviously, the overall work is

Wtotal = W12 + W21 = 0. (11.8)

Thus, the machine does not release net work.

To enable the machine to release net work, expansion must take place at a higher
pressure level than compression. This can be achieved by supplying thermal energy
after compression, see Fig. 11.6b. While heat is supplied, the position of the piston
is fixed, i.e. an isochoric change of state takes place. Consequently, state (3) has
the same volume as state (2), but both temperature and pressure have increased.
The subsequent expansion to state (4) now takes place at a higher pressure level.
Therefore, the work released is greater than the work supplied during compression:

|W34 | > W12 . (11.9)

If the gas is compressed again in state (4), state (3) is reached again. Assuming
that the heat supply can still be isochoric,5 the temperature and pressure of the gas
continue to rise. At this higher pressure, expansion can be initiated so that work
is released. If these steps are repeated periodically, temperature and pressure at the
bottom dead centre (BDC) will continue to rise until the machine overheats. Further-
more, since expansion and compression in such a machine follow the same curve, the
machine also does not release any work. The heat supplied only leads to an increase
in temperature, see Fig. 11.7.
To operate the machine continuously, thermal energy must be released in state
(4), e.g. by isochoric cooling, cf. Fig. 11.6b. Thus, the temperature decreases and
the initial state (1) can be reached again. Once back in state (1), the entire process
can restart so that the machine can be permanently operated without the machine
overheating. According to Fig. 11.6b work is released from (3) → (4), whereas work
is supplied from (1) → (2). The net work can be illustrated in a p, V -diagram by the
grey enclosed area. Obviously, the thermodynamic cycle consists of four successive

4 Under these premises, the entire work consists of pure volume work. Areas in a p, V -diagram
represent volume work or pressure work in open systems.
5 Of course, the heat supply depends on the temperature level of the environment.
11.2 Closed Systems 163

Fig. 11.7 Closed


system—no cooling

changes of state. The first law of thermodynamics can be applied to any change of
state6 :

• Change of state (1) to (2)

Q 12 + W12 = U2 − U1 (11.10)

• Change of state (2) to (3)

Q 23 + W23 = U3 − U2 (11.11)

=0

• Change of state (3) to (4)

Q 34 + W34 = U4 − U3 (11.12)

• Change of state (4) to (1)

Q 41 + W41 = U1 − U4 (11.13)

=0

6 Changes in kinetic as well as potential energy are ignored, see Sect. 11.2.2. However, if these
effects are not neglected, e.g. in a vertical cylinder, the conclusion would be the same, since they
disappear anyway by the addition of all successive changes of state.
164 11 First Law of Thermodynamics

A summation of Eqs. 11.10 to 11.13, yields


 
Q+ W =0 (11.14)
  
Weff

The state values U1 , . . . , U4 disappear, since the initial state (1) is reached again
after several state changes. Equation 11.14 can be applied to any thermodynamic
cycle. Equation 11.14 further proves that the net work is equal to the sum of the
thermal energies. In the case of the proposed machine, the net work released, i.e. the
grey enclosed area7 in Fig. 11.6b, is the difference between heat supplied and heat
released, since the heat released is negative:

Weff = −Q Heating − Q Cooling < 0 (11.15)

Note that Q heating > |Q cooling |, so that the net work is negative, i.e. from the system’s
point of view, work is released.
Equation 11.14 can therefore be applied to describe the energy conversion from
thermal to mechanical energy.

11.2.2 Systems in Motion

Now the focus is on closed systems that are in motion or change their position in a
gravitational field. The total energy of such a system is shown in Fig. 11.8. Obviously,
a system does not only possess internal energy U , but also potential energy E pot and
kinetic energy E kin . Both, potential and kinetic energy, are also summarised as outer
energies8 :

1  
E = U + E kin + E pot = (U − U0 ) + m c2 − c02 + mg(z − z 0 ). (11.16)
2
Note that a reference level is required for the specification of energies. This is con-
sidered in Eq. 11.16 by reference levels for the internal energy U0 , the velocity c0 and
the vertical position z 0 . According to the law of conservation of energy, a change in
an internal energetic state is associated with an exchange of energy across the system
boundary. Hence, it follows:

E 2 − E 1 = Q 12 + W12 . (11.17)

7 Note that the p, V -diagram represents the volume work. The volume work is identical to the total

work if there is no dissipation and changes in the outer energies are ignored.
8 However, if the centre of gravity is fixed, neither the potential nor the kinetic energy changes. As

a result, they are ignored in Eq. 11.6 for systems at rest.


11.2 Closed Systems 165

Fig. 11.8 Closed system in


motion

The substitution of the total energy by Eq. 11.16 leads to a general notation of the
first law of thermodynamics for closed systems:

1  
U2 − U1 + m · c22 − c12 + mg · (z 2 − z 1 ) = Q 12 + W12 (11.18)
2

Since two states are being compared, the reference level obviously disappears. Con-
sequently, during a change of state of (1) → (2) the reference level must not be
changed. By dividing by the mass of the system, the specific notation of the first law
of thermodynamics is obtained:

1 2 
u2 − u1 + c2 − c12 + g · (z 2 − z 1 ) = q12 + w12 . (11.19)
2

Furthermore, in differential notation9 it is

du + c dc + g dz = δq + δw (11.20)
  
dea

The first law of thermodynamics for closed systems can be motivated with the fol-
lowing Fig. 11.9. Energy is represented by bubbles so that the total energy of a system
in states (1) and (2) can be easily counted. Obviously, the change in the total energy
of the system, i.e. the difference of the bubbles in the system in the states (2) and (1),
is balanced by the energies W12 and Q 12 , which both cross the system boundary. If
energy is supplied from outside, i.e. if it is positive, the number of bubbles increases.
If the external energies are negative, which means that the system is releasing energy,
the number of bubbles in state (2) is less than in state (1). This correlation is given
by Eqs. 11.17 respectively 11.18. The left-hand side of these equations indicates the

9 Integration of Eq. 11.20 leads to Eq. 11.19.


166 11 First Law of Thermodynamics

Fig. 11.9 First law of thermodynamics for closed systems, see Sect. 3.2.5

energy change, while the right-hand side indicates the cause of this internal change,
i.e. the external energetic influences.

11.2.3 Partial Energy Equation

Figure 11.10 illustrates the procedure for the energetic balancing of closed systems.
It shows an ideal gas in a cylinder that is closed by a movable piston. However, it is
advisable and recommended to follow a clear, structured approach:
1. Formulation of the first law of thermodynamics:
The energy balance, see Eq. 11.18, takes into account all external influences, i.e.
both thermal energy Q 12 and non-thermal energy W12 :

1  
U2 − U1 + m · c22 − c12 + mg · (z 2 − z 1 ) = Q 12 + W12 (11.21)
2

Fig. 11.10 Partial energy equation for closed systems


11.2 Closed Systems 167

Obviously, according to Fig. 11.10, the external energies cause a change of the
internal state, i.e. the internal energy, as well as of the outer state such as the
kinetic and potential energy, which are related to the centre of gravity.
2. Formulation of the partial energy equation for the exchanged work:
In this second step, it should be specified how the non-thermal energy W12 is
composed. This is achieved by applying the so-called partial energy equation,
which always has to be adapted to the problem at hand. The underlying question
should be: What kind of work is part of W12 ? A generic approach is given by
Eq. 11.22:
W12 = WV,12 + WMech,12 + 12 (11.22)

In this case, i.e. in Fig. 11.10, the total non-thermal10 energy W12 consists of vol-
ume work WV,12 , since the gas is compressed, and additional mechanical work
WMech,12 to lift and accelerate the centre of gravity. Furthermore, as the sketched
turbulent vortices show, the dissipation 12 consumes additional energy: For
sustaining the turbulence, respectively dissipation, energy is required. More pre-
cisely, this energy is part of the work W12 . If a system releases work W12 , dissipa-
tion reduces this work. In the opposite case, when work is supplied to a system, the
amount of work required is greater than in the reversible case in order to achieve
the same result.11 It is important to mention that the partial energy equation is not
an all-inclusive energy equation, as it only focuses on the non-thermal energies.
Hence, no thermal energies are covered by this equation.
3. Formulation of the effective work:
If the open system is operated in an environment with penv > 0, the work released
by the gas during expansion partially compresses the environment. If, in contrast,
the gas is compressed, the expanding environment supports the compression, see
Fig. 11.11. According to Fig. 11.11, the following work is effectively transferred
at the piston rod, see also Sect. 9.2.3:

Weff = W12 + W12,env = W12 + penv (V2 − V1 ) . (11.23)

However, the following two cases can occur:


– Compression, i.e. V2 < V1 :

Weff = W12 + W12,env = W12 + penv (V2 − V1 ) . (11.24)


  
<0

10 Volume work, dissipation and work related to the centre of gravity are in the end mechanical
work. Thus, the focus is on such mechanical work.
11 This is because dissipation always consumes energy. From experience, e.g. with an air pump, it

is known that with friction a greater effort is required to achieve the same result, e.g. compression
to the same volume, than if the process is reversible, i.e. free of dissipation.
168 11 First Law of Thermodynamics

Fig. 11.11 Effective work

As an example:
=W12,env

Weff = 8 J = +10
 J −2 J . (11.25)
=W12

Less work than W12 is required at the piston rod for the compression as the
environment expands and supports the compression.

– Expansion, i.e. V2 > V1 :

Weff = W12 + W12,env = W12 + penv (V2 − V1 ) . (11.26)


  
>0

As an example:
=W12,env

Weff = −8 J = −10
  +2 J .
J (11.27)
=W12

Less work than W12 can be utilised by the piston rod, as work is needed to
compress the environment.

Problem 11.1 Please state the first law of thermodynamics and the partial energy
equation for the three cases given in Fig. 11.12!

Solution

(a) In this case the first law of thermodynamics reads as

1  
W12 + Q 12 = U2 − U1 + m c22 − c12 + mg (z 2 − z 1 ) (11.28)
2
11.2 Closed Systems 169

Fig. 11.12 a Supply of electrical energy in an isochoric system, b Supply of electrical energy in a
non-isochoric system, c Supply of electrical energy with an electrical capacitor

Since the centre of gravity does not change, only the inner state is relevant.
Furthermore, in this case, no change in the electrical state of the system is taken
into account. Consequently, the first law of thermodynamics simplifies to

W12 + Q 12 = U2 − U1 . (11.29)

The partial energy equation for that case follows12



W12 = U I dt = WV,12 + WMech,12 +12 . (11.30)
     
t =0 =0

This equation expresses that the entire work W12 is supplied electrically. No
mechanical work is transferred because the centre of gravity is fixed. In addition,
the volume is constant in this case, so no volume work is transferred either.
According to
W12 = U I dt = 12 (11.31)
t

the entire electrical energy is dissipated.13 This is the cause for an increase of
temperature. However, this process can never be reversible, since the supplied
energy can only be partially converted back, see Chap. 16.

12 U denotes the voltage potential.


13 Although the fluid is not moving in this isochoric case, and therefore there are no turbulence
eddies, dissipation occurs in the system. This is due to the dissipated energy at the electrical resis-
tance.
170 11 First Law of Thermodynamics

(b) Starting with the first law of thermodynamics for the second case

1  
W12 + Q 12 = U2 − U1 + m c22 − c12 + mg (z 2 − z 1 ) (11.32)
2
Assuming that the system is at rest in states (1) and (2), the vertical position of
the centre of gravity has changed. Analogous to (a), the electrical state of the
system is not considered. Thus, the first law of thermodynamics obeys

W12 + Q 12 = U2 − U1 + mg (z 2 − z 1 ) . (11.33)

For this case, the partial energy equation follows:



W12 = U I dt = WV,12 + WMech,12 +12 . (11.34)
  
t =mg(z 2 −z 1 )

In this case, dissipation, mechanical as well as volume work occur, see


Fig. 11.12b.
(c) This last case represents a rather specific example, as the electrical state of the
system is also to be considered. This is indicated by the symbol for a plate
capacitor within the system boundary. An electrical capacitor is able to store
electrical energy according to14

1
E el,12 = C(U22 − U12 ) = Wel,12 . (11.35)
2
Therefore, the first law of thermodynamics for this case, assuming states (1) and
(2) to be at rest, is as follows15

W12 + Q 12 = U2 − U1 + mg (z 2 − z 1 ) + E el,12 . (11.36)

However, for this case the partial energy equation reads as16

W12 = U I dt = WV,12 + WMech,12 + Wel,12 +12 . (11.37)
     
t =mg(z 2 −z 1 ) =E el,12

Combining Eqs. 11.36 and 11.37 yields

Q 12 + WV,12 + 12 = U2 − U1 . (11.38)

14 Once again, in this equation U denotes the voltage potential and not the internal energy U .
15 The change in the electrical state must now be taken into account on the right-hand side of the
first law of thermodynamics.
16 The work that varies the electrical state of the system must now be additional considered. However,

the work crossing the system boundary equals the electrical work.
11.2 Closed Systems 171

Fig. 11.13 a Reversible, adiabatic compression, b Irreversible, adiabatic compression

Example: Reversible Versus Irreversible

The first law of thermodynamics can be applied to all changes in a closed sys-
tem, regardless of whether the change of state is reversible or irreversible. However,
let us investigate these two case, see Fig. 11.13. Figure 11.13a shows an adiabatic,
reversible compression, while Fig. 11.13b represents an adiabatic, irreversible com-
pression. The fluid inside the cylinder is supposed to be an ideal gas.
• Reversible, adiabatic (1) → (2), Fig. 11.13a

The first law of thermodynamics for this case reads as17 :

Q 12 +W12 = U2 − U1 . (11.39)

=0

Since the work W12 consists purely of volume work,18 the partial energy equation
follows:
W12 = WV,12 > 0. (11.40)

Consequently, the work supplied increases the internal energy of the system, i.e.

WV,12 = U2 − U1 > 0. (11.41)

With the caloric equation of state, that is introduced in Chap. 12, it follows that

U2 − U1 = mcv (T2 − T1 ) > 0. (11.42)

Thus the temperature rises from (1) to (2):

T2 > T1 . (11.43)

17 Change of outer energies can be ignored, since the system is at rest in states (1) and (2) and the
vertical position of the centre of gravity does not vary.
18 No dissipation, no mechanical work due to horizontal cylinder.
172 11 First Law of Thermodynamics

• Irreversible, adiabatic (1) → (2 ), Fig. 11.13b

In this case, the first law of thermodynamics has the same structure as in the
reversible case. The dissipation therefore does not occur directly:

Q 12 +W12 = U2 − U1 . (11.44)



=0

However, the partial energy equation includes the dissipation 12 :

W12 = WV,12’ + 12 . (11.45)

The work W12 that passes the system boundary is used both for compression and
for supporting dissipation.19 This finally leads to

WV,12’ + 12 = U2 − U1 . (11.46)

Both WV,12’ and 12 are positive, so that

U2 − U1 = mcv (T2 − T1 ) > 0. (11.47)

Again, from (1) → (2 ) temperature increases.


Experience shows that irreversible changes of state require more work than reversible
changes of state to achieve the same effect, i.e.

W12 > W12 . (11.48)

Hence, it follows that U2 > U2 . Applying the caloric equation of state, see Chap. 12,
leads to T2 > T2 . Both changes of state can be illustrated in a p, V -diagram, see
Fig. 11.14a. As indicated, reversible and irreversible compression reach the same
final volume, i.e. both achieve the same effect. Since temperature T2 is greater than
T2 , state (2 ) lies vertically above (2). Note that the area under a change of state
in a p, V -diagram symbolises the volume work. Consequently, in the case of an
irreversible change of state, the greater temperature leads to a greater pressure, so
that the volume work of (1) → (2 ) is greater than the volume work (1) → (2):

WV,12’ > WV,12 (11.49)

Note that the dissipation is not visible in the diagram. However, the total work yields

>0

WV,12’ + 12 > WV,12 = W12 (11.50)
  
=W12

19 The dissipation always consumes energy. This energy has to be supplied.


11.2 Closed Systems 173

Fig. 11.14 Partial energy equation for closed systems

so that
W12 > W12 . (11.51)

With the second law of thermodynamics, see Chap. 13, the entire process can
be explained in detail.20 Furthermore, the changes of state can be stated in a T, s-
diagram, see Fig. 11.14b. However, the T, s-diagram is introduced in Sect. 13.4.
However, for the entropy, cf. Sect. 12.2, it follows:

δq δψ
ds = + . (11.52)
T T

An adiabatic, reversible change of state is isentropic.21 Obviously, the entropy


remains constant from (1) → (2). Chapter 13 shows that the area under a change
of state in a T, s-diagram is the sum of specific heat q and specific dissipation ψ.
According to Eq. 11.52 the entropy rises from (1) → (2 ). In order to reach the same
final volume as in the reversible case, i.e. the same specific volume, since the mass is
constant, state (2 ) lies to the right and above of state (2). Obviously, temperature T2
is greater than T2 . Therefore, both p, V - and T, s-diagram lead to the same result.
Example 11.2 Let us apply the principles of the first law of thermodynamics to the
following example of a horizontal cylinder filled with an ideal gas, cf. Fig. 11.15.
• Case 1 (frictionless, adiabatic) (1) → (2)
The first law of thermodynamics obeys

W12 + Q 12 = U2 − U1 . (11.53)
 
>0 =0

20 Thus, it is not only necessary to rely on the experience that systems with friction require more
effort than systems without friction.
21 This is discussed in Sect. 13.5!.
174 11 First Law of Thermodynamics

Fig. 11.15 Partial energy equation for closed systems

Since it is an ideal gas, the caloric equation of state is applied, so that for the
temperature in state (2) is
W12
T2 = T1 + . (11.54)
mcv

• Case 2 (friction, adiabatic, same entire work as in case 1: W12 ) (1) → (3)
The same temperature is reached because the same work is supplied to the adiabatic
system as in case 1:
W13 = W12 = mcv (T3 − T1 ) . (11.55)

Consequently, it is
T2 = T3 . (11.56)

However, the supplied work W13 is composed of volume work, i.e. the area under
the curve in the p, V -diagram, and dissipation. Hence, the partial energy equation
11.2 Closed Systems 175

reads as:
3
W13 = W12 = − p dV +13 . (11.57)
1
  
WV,13

Obviously, the volume work is the total work reduced by the dissipation 13 , i.e.

WV,13 = W12 − 13 . (11.58)

State (3) can be fixed in a p, V -diagram, see Fig. 11.15. States ((2) and (3) must
be on the same isotherm. However, the area beneath (1) → (2), i.e. the volume
work (1) → (2), is greater than the area beneath (1) → (3), i.e. the volume work
(1) → (3). Thus, with the same amount of work, the same final volume cannot be
achieved with an irreversible change of state as with a reversible process.
• Case 3 (Friction, cooled) (1) → (2 ) → (2)

In order to reach the same final volume V2 in case of irreversible operation, the
amount of work exceeds22 W12 :

W12 > W12 . (11.59)

The first step (1) → (2 ) is done adiabatically. The first law of thermodynamics
follows according to
W12 = U2 − U1 > U2 − U1 . (11.60)

Hence, temperature T2 is greater than temperature T2 . The partial energy equations
obeys
W12 = WV,12’ + 12 > W12 . (11.61)

Now that the compression is completed to V2 , the temperature must be adjusted to


reach state (2). The change of state (2 ) → (2) is therefore conducted isochorically,
i.e. without transfer of work. Heat must be released so that temperature drops.
According to the first law of thermodynamics it is

W2 2 +Q 2 2 = U2 − U2 < 0. (11.62)



=0

It obviously follows that

Q 2 2 = U2 − U2 = W12 − W12 . (11.63)

22 As case 2 has shown, the same amount of work in reversible/irreversible operation does not result
in the same compression rate. In irreversible operation, compression stops before reaching the end
position. Obviously, to achieve the same compression ratio, more work has to be supplied.
176 11 First Law of Thermodynamics

Combining with the partial energy equation yields

Q 2 2 = WV,12 − WV,12’ + 12 − 12 . (11.64)

Thus, the hatched area in Fig. 11.15 represents

WV,12’ − WV,12 = −Q 2 2 − 12 (11.65)

Problem 11.2 Two identical vertical pipes with a diameter of D = 0.1 m are con-
nected by a thin tube and a valve, see Problem 2.1. One pipe is filled with water
up to a height of H = 10 m. Water has a density of ρ = 1000 mkg3 , the other tube is
initially empty. Now the valve is opened. After some time, an equilibrium of the
water quantity is reached. By what absolute value does the internal energy of the
water change, if the system does not exchange any energy with the environment?

Solution

Applying the first law of thermodynamics for closed systems reads as, see Fig. 11.16:

Q +W = U2 − U1 + E pot + E kin . (11.66)


 12  12      
external impact change of internal state change of outer state

Since there is no energetic exchange with the environment, it follows that Q 12 = 0


and W12 = 0. In addition, the system is at rest in states (1) and (2), so that the kinetic
energy does not change. Hence, Eq. 11.66 simplifies to:

0 = U2 − U1 + E pot . (11.67)

Fig. 11.16 Solution Problem 11.2


11.2 Closed Systems 177

The solution of the Problem 2.1 has shown that the centre of gravity is relevant for
the calculation of the change in potential energy:

U = U2 − U1 = −E pot = −mg (z 2 − z 1 ) . (11.68)

This finally results in




H H
U = −mg − = +1926.2 J (11.69)
4 2

While the potential energy of the system decreases, its internal energy increases so
that the total energy remains constant. Let us now apply the partial energy equation.
Since the entire system does not exchange work with the environment,23 it reads as:

W12 = 0 = WV,12 + WMech,12 +12 (11.70)


     
=0 =mg(z 2 −z 1 )

respectively

H H
12 = −mg − = +1926.2 J. (11.71)
4 2

The total mechanical energy is finally dissipated.

Problem 11.3 A horizontal cylinder, closed by a movable piston, contains 20 dm3


of an ideal gas under a pressure of 1 bar. The gas is compressed by a piston to a
pressure of 5 bar. The dissipated energy is 10 % of the total energy feed. The change
of state can be approximated by p · V 1.25 = const. for the entire change of state.
(a) How far does the piston have to be pushed into the cylinder? The diameter of
the cylinder is d = 20 cm.
(b) How much work has to be supplied to the system?
(c) How much heat must be transferred across the cylinder walls if the internal
energy of the gas increases by 1 kJ during compression?

Solution

(a) The change of state follows p · V 1.25 = const., so this equation can be applied
in state (1) and in state (2):

23 The environment does not expand and is not compressed, since the volume of the water remains
constant all time: Although one column compresses the environment, the work applied to do so is fed
back into the system at the other water/environment interface, where expansion of the environment
takes place. Compression work and expansion work at the environment ( penv = const.) are identical
in absolute value.
178 11 First Law of Thermodynamics

p1 V11.25 = p2 V21.25 . (11.72)

This equation can be solved for the volume V2 in state (2):



p1
V2 = V1 · 1.25 = 5.52 dm3 . (11.73)
p2

Thus, the change of volume yields

V = |V2 − V1 | = 14.48 dm3 . (11.74)

Following geometrical considerations:


π 2
V = d s. (11.75)
4
Therefore, the distance that the piston has to be moved is

4 V
s = = 4.61 dm (11.76)
π d2

(b) The first law of thermodynamics applied to this problem obeys

W12 + Q 12 = U2 − U1 . (11.77)

The dissipation 12 does not occur directly since energy is dissipated within the
system. However, the partial energy equation reads as:

W12 = W12,V + 12 . (11.78)

It is known that 12 = 0.1W12 . Consequently one obtains

W12 = W12,V + 0.1W12 . (11.79)

Solving for W12 :


W12 = 1.11W12,V . (11.80)

The volume work can be calculated according to

2
W12,V = − p dV (11.81)
1

Thus, it is
2
W12 = −1.11 p dV. (11.82)
1
11.2 Closed Systems 179

To solve the integral, the direction of the change of state is needed. However,
that description is given as p · V 1.25 = const. In Chap. 12 this change of state is
introduced as polytropic change of state. Applied to this problem it states

1.25
V1
p = p (V ) = p1 . (11.83)
V

Thus, Eq. 11.82 simplifies to

2
1.25 2
V1
W12 = −1.11 p1 dV = −1.11 p1 V11.25 V −1.25 dV. (11.84)
V
1 1

The solution finally is



W12 = 4.44 p1 V11.25 V2−0.25 − V1−0.25 = 3.3714 × 103 J (11.85)

(c) Applying the first law of thermodynamics yields

Q 12 + W12 = U2 − U1 . (11.86)

Thus, it follows

Q 12 = U2 − U1 − W12 = −2.372 × 103 J (11.87)

Obviously, the system needs to be cooled.

Problem 11.4 A vertical cylinder with an inner diameter of 250 mm is filled with
a gaseous mixture that is in thermodynamic equilibrium with the environment. The
cylinder is closed against the environment by a gas-tight, frictionless moving piston.
The weight of the piston as well as ambient pressure lead to a gas pressure of 4 bar
in the cylinder. Now the mixture is ignited and a combustion takes place. Hence,
the piston moves upwards and heat is exchanged with the environment. The entire
change of state runs so slowly, that the mixture is under a constant pressure of 4 bar.
Finally, thermodynamic equilibrium is reached. By what absolute value does the
internal energy of the gas change, if the gas releases 10 kJ of heat and the piston is
lifted by 75 mm? Ambient pressure is penv = 1 bar. Potential energy of the gas can
be ignored. Derive the energy balances for the following systems:
(I) Cylinder content on its own
(II) Cylinder content and piston
180 11 First Law of Thermodynamics

Fig. 11.17 Solution to Problem 11.4—Isobaric change of state

Solution

The solution is derived in general terms, i.e. dissipation in the gas is taken into
account in order to clarify how it is to be treated. However, since the change of state
in this particular case is slow and any acceleration of the piston can be neglected,24
no dissipation occurs, see Example 7.6.

Part (I)

System boundary: see Fig. 11.17a

• First law of thermodynamics:

W12,G + Q 12,G + Q Z = UG + E Pot,G (11.88)


     
Q total =−10kJ =0

respectively
W12,G + Q total = UG . (11.89)

• With the partial energy equation according to Eq. 11.22:

W12,G = W12,G,V + 12,G + W12,G,mech . (11.90)


  
=0

Since the change of state runs very slowly,25 there is no dissipation, i.e. 12,G = 0,
see Theorem 7.16:

24 A quasi-static change of state is assumed.


25 As mentioned explicitly in the problem description.
11.2 Closed Systems 181

2
W12,G = W12,G,V + 12,G = − pG dV = − pG A z. (11.91)
  
=0 1

Combining Eqs. 11.89 and 11.91 finally yields

UG = − pG A z + Q total = −11.47kJ (11.92)

Part (II)

System boundary: see Fig. 11.17b


• First law of thermodynamics:

Q 12,G + Q 12,P + W12,env = UG + UP + E Pot,P + E Pot,G (11.93)


  
=0

respectively

Q 12,G + Q 12,P + W12,env = UG + UP + m P g z. (11.94)

• The partial energy equation for W12,env can be given in two notations, since W12,env
appears at the interface of system and environment:
1. From the system’s point of view, according to the partial energy Eq. 11.22, it
is

W12,env = W12,G,V + 12,G +W12,P,mech = − pG Az + m P g z. (11.95)


  
=0

Obviously, from this perspective work is required for the volume change of
the gas and for lifting the piston.26 These two mechanisms are part of the
system boundary. As already mentioned, there is no dissipation within the gas.
Combining Eqs. 11.94 and 11.95 leads to

Q 12,G + Q 12,P − pG A z = UG + UP (11.96)

2. From the environmental’s point of view the work W12,env is purely required for
compressing the environment. 12,env does not appear because the environment
is always considered homogeneous,27 E a,env does not appear because the
environment is huge and at rest:

26Work to lift the centre of gravity of the gas is negligible according to the problem description.
27Homogeneous means that there are no imperfections. Consequently, there is no dissipation, see
Sect. 14.4 for details.
182 11 First Law of Thermodynamics

2
W12,env = W12,env,V = penv dVenv = penv Venv . (11.97)
1

In this case there is no minus sign in front of the integral, because according
to Fig. 11.17b the arrow of W12,env points out of the environment, i.e. contrary
to the sign convention, see Fig. 9.4. However, as the volume of the system
increases, the volume of the environment decreases and vice versa, so that

Venv = −V = − penv A z. (11.98)

Hence, it is
W12,env = − penv A z. (11.99)

Combining Eqs. 11.94 and 11.99 leads to

UG = Q 12,G + Q 12,P − penv A z − m P g z − UP (11.100)

What is the correlation between parts I) and II)? The solution must be independent of
the choice of system boundary. To clarify this, a third system boundary is examined,
see Fig. 11.17c.
• The first law of thermodynamics for the piston obeys

W12,env − W12,G + Q 12,P − Q Z = UP + E Pot,P . (11.101)


  
m P g z

Substitution of the transferred works by Eqs. 11.91 and 11.99 leads to

− penv A z + pG A z + Q 12,P − Q Z = UP + m P g z (11.102)

respectively

0 = UP + m P g z + penv A z − pG A z − Q 12,P + Q Z . (11.103)

• Summation of Eqs. 11.100 and 11.103 brings

UG = − pG A z + Q total (11.104)

Consequently, Eqs. 11.92 and 11.104 are identical. Consequently, it does not matter
which system boundary is chosen: The result must always be boundary-independent.
Example 7.6 has shown that a balance of forces, see also Eq. 4.5, according to

mPg
pG = penv + (11.105)
A
11.2 Closed Systems 183

Fig. 11.18 Solution to


Problem 11.5

can be applied for slow changes of state, a non-accelerated and frictionless piston
and for neglecting the kinetic and potential energy of the gas. Thus, Eq. 11.103
simplifies to
UP = Q 12,P − Q Z (11.106)

The internal energy of the piston is determined solely28 by the net heat. In this
particular case, since the system is at thermal equilibrium with the environment in
states (1) and (2), the change of the piston’s internal energy is zero,29 i.e.

UP = 0 (11.107)

However, the change in the internal energy of the gas, although the temperatures T1
and T2 are identical, is not zero because the chemically bonded energy of the gas
was released during the combustion process.30

Problem 11.5 Problem 11.2 is now to be studied in more detail. The change of state
starts in the initial position (1), where one column is filled with water while the other
is empty. Derive an equation, that shows the level of water as a function of time,
see Fig. 11.18. Consider a reversible as well as an irreversible change of state. The
system still is supposed to be adiabatic.

28 As long as the piston itself moves frictionless, see Example 7.6.


29 Since the temperature of the piston does not change. It has ambient temperature in states (1) and
(2).
30 In fact, the gas from state (1) to state (2) has been replaced, i.e. its chemical composition has

changed.
184 11 First Law of Thermodynamics

Solution

First, the partial energy equation for work is applied. Since there is no work exchanged
with the environment,31 it reads as:

W12 = 0 =  + WV,12 +Wmech,12 . (11.108)


  
=0

Since the volume is constant, there is no volume work. In differential notation, the
partial energy equation follows:

δ + m total c dc + m total g dz .
0 =  (11.109)
     
Dissipation kinetic potential

• With regard to dissipation, a mathematical model is required. From fluid dynam-


ics32 it is known that the dissipation in a laminar flow is proportional to the distance
travelled and the velocity of the fluid. The distance is the path along which the
liquid column is moved, i.e. |dy|, so that

dy
δ ∝ | | |dy|. (11.110)
dt

At this point it is important to keep the dissipation positive, as dy as well as dy


dt
can be positive and negative. Introducing a proportional constant ξ it follows that

dy
δ = ξ | | |dy|. (11.111)
dt
dy
However, dy and dt
always have the same sign, so that the product is always
positive, i.e.
dy
δ = ξ dy. (11.112)
dt
• The kinetic energy obeys

dy dc d2 y
m total c dc = m total dc = m total dy = m total 2 dy. (11.113)
dt dt dt

31 The environment does not expand and is not compressed, since the volume of the water remains
constant all time: Although one column compresses the environment, the work applied to do so is fed
back into the system at the other water/environment interface, where expansion of the environment
takes place. Compression work and expansion work at the environment ( penv = const.) are identical
in absolute value.
32 For a laminar flow the pressure drop is p = ρ c2 λ l with λ = 64η . The combination brings
2 d ρcd
32uηl p 32uηl
p = d2
. The specific dissipation is ψ = ρ = ρd 2
, the entire dissipation is  = ψm total .
11.2 Closed Systems 185

Since the velocity of the entire mass is equal to dy


dt
, the mass has the same velocity
as the water surface.
• Now the position z of the entire centre of gravity must be calculated as a function
of y. The mass of the left column is


H
m left = ρ A +y . (11.114)
2

The vertical position of its centre of gravity is




1 H
z left = +y . (11.115)
2 2

The mass in the right column follows




H
m right = ρ A −y . (11.116)
2

The vertical position of its centre of gravity is




1 H
z right = −y . (11.117)
2 2

In order to find the vertical position z of the entire centre of gravity, the following
equation can be applied33

z left m left + z right m right H y2


z= = + . (11.118)
m left + m right 4 H

Thus, its differential dz is




y2 1  2  2y
dz = d = d y = dy. (11.119)
H H H

Hence, the partial energy Eq. 11.109 simplifies to:

dy d2 y 2y
0=ξ dy + m total 2 dy + m total g dy. (11.120)
dt dt H
Dividing by the mass of the system m total and dividing by the change of position of
the water surface dy lead to:

33 As known form mechanics.


186 11 First Law of Thermodynamics

Fig. 11.19 Numerical solution to Problem 11.5

d2 y ξ dy 2g
2
+ + y=0 (11.121)
dt m total dt H

A numerical solution is given in Fig. 11.19. Obviously, it is dissipation that dampens


the system. A function for the temporal temperature development can be derived by
a comparison of the first law of thermodynamics

δW + δ Q = 0 = dU +m total g dz + m total c dc
 (11.122)
m total cW dT

and the partial energy equation

δW = 0 = δWV +δ + m total g dz + m total c dc. (11.123)



=0

Obviously, it is
δ = dU = mcW dT. (11.124)

The dissipation can be substituted by Eq. 11.112, so that

dy
ξ dy = m total cW dT. (11.125)
dt
Rearranging leads to
11.2 Closed Systems 187

ξ dy
dT = dy. (11.126)
m total cW dt

Dividing by dt brings
dT ξ dy dy
= . (11.127)
dt m total cW dt dt

Finally it is

2
dT ξ dy
= >0 (11.128)
dt m total cW dt

A numerical solution is illustrated in Fig. 11.19 as well.

11.3 Open Systems

An example for an open system is illustrated in Fig. 11.20. It shows two components
connected in series: A gas is compressed in a compressor and cooled in a subsequent
heat exchanger. However, open systems are characterised by an inflow respectively
outflow, i.e. mass passes a system boundary. In many technical applications, e.g.
a compressor station, changes of state do not take place as a batch process, i.e. in
discontinuous operation, but continuously.34 In the example illustrated, a gas flow
enters the system indicated by the dashed boundary. The compressor increases the
pressure of the fluid while its temperature rises as well. This compression requires
technical work that exceeds the selected system boundary. In the subsequent heat
exchanger, the fluid is cooled, i.e. heat is released into the environment. Here, as
well, energy is thus released across the system boundary. According to Fig. 11.20
the system boundary is permeable to a heat flux in the cooler. In addition, there may
be insulation losses in the compressor or in the connecting pipes. The system is also
open to mechanical energy supplied to the compressor. However, the open system
shown can be simplified by an equivalent model, see Fig. 11.21. For the inlet (1), a
mass flow ṁ 1 and a corresponding velocity c1 are given, i.e.

1
ṁ 1 = ρ1 V̇1 = A1 c1 . (11.129)
v1

In addition, the vertical position in the gravitational field is represented by z 1 . The


outlet (2) follows accordingly. In case the system is in a stationary state, the following
results, known as equation of continuity:

ṁ 1 = ṁ 2 . (11.130)

34 Heat is not only transferred once, but continuously. Therefore, a heat flux instead of a quantity
of heat is relevant.
188 11 First Law of Thermodynamics

Fig. 11.20 Compression process

Fig. 11.21 Schematic diagram (equivalent model)

Thus, the volume flows in steady state obey

V̇1 ρ2 v1 p2 T1
= = = . (11.131)
V̇2 ρ1 v2 p1 T2

In addition to the mass flows, both heat and technical work35 pass the system bound-
ary. It is important to emphasise that the heat flux Q̇ 12 includes the entire heat passing
the system boundary from (1) to (2), i.e. also heat conduction within the fluid at the
inlet and outlet is part of Q̇ 12 .

11.3.1 Formulation of the First Law of Thermodynamics


for Open Systems

The principles that have been derived for closed systems in the previous section are
now applied to open systems. At t = 0 the mass inflow is achieved by a piston that

35 Technical work is always related to open systems.


11.3 Open Systems 189

Fig. 11.22 From closed to open systems

Fig. 11.23 Energy balance open system

supplies shifting work to a piece of mass m 1 at the inlet, cf. Fig. 11.22. Within a
period of time t, both technical work Wt,t and heat Q t are supplied. After the
regarded time step t, the outflow in this equivalent model is realised by a second
piston that delivers the required shifting work. The system boundary is therefore
mass-permeable. In Fig. 11.23 the extensive state value energy is represented by
bubbles: Energy is within the system and also crosses the system boundary. As
already known, energy is conserved, i.e. it can neither be generated nor destroyed.36
Consequently, the sum of the total energy passing the system boundary is equal to
the energy change E Sys that can be observed within the system, cf. also Eq. 3.12:

36 Thus, there is no source/sink term for energy.


190 11 First Law of Thermodynamics

Fig. 11.24 Energy within an open system

 
E Sys,t − E Sys,0 = E in,i − E out, j (11.132)
i j

E Sys,0 represents the total energy of the system at the time t = 0. E Sys,t is the
total energy of the system at the time t = t. The term E Sys,t − E Sys,0 therefore
states by what amount the energy within the system has changed in a period of time
t. Such a change of energy  within the system is caused by the flow across the
system boundary: Inflows, i.e. E in,i , increase the energy of the system, outflows,
 i
i.e. E out, j , decrease the energy of the system. In other words, the energy of the
j
system at t = t is equal to the initial energy plus the energy inflows minus the
energy outflows, i.e.
 
E Sys,t = E Sys,0 + E in,i − E out, j . (11.133)
i j

The next section clarifies how these energies, both within the system and across the
system boundary, are composed.

Energy content of the system

The energy content of the system is illustrated in Fig. 11.24. It is important that any
energy balance must be carried out extensively,37 i.e. the mass38 of the system must

37 Energy is constant, not specific energy.


38 Mass as an extensive state value represents the size of a system.
11.3 Open Systems 191

be known as well.39 Let the mass at t = 0 be m Sys,0 and the mass at t = t be m Sys,t .
The system contains internal energy USys at both times, cf. Sect. 2.4.2. Furthermore,
it possesses kinetic and potential energy, due to cSys respectively z Sys with respect to
the centre of gravity. Obviously, at the times t = 0 and t = t, the system can be
considered as a closed system whose energy content can be easily calculated, i.e.
 
1 2
E Sys,0 = m Sys,0 · u Sys,0 + cSys,0 + g · z Sys,0 (11.134)
2

and  
1 2
E Sys,t = m Sys,t · u Sys,t + cSys,t + g · z Sys,t . (11.135)
2

However, since mass is conserved, see Fig. 11.25, the overall mass balance obeys
 
m Sys,t = m Sys,0 + m in,i − m out, j . (11.136)
i j

The mass at t = t is the initial mass at t = 0 increased by the inflow mass and
decreased by the outflow mass, cf. Eq. 3.12. In another notation it reads as
 
m Sys,t − m Sys,0 = m in,i − m out, j . (11.137)
i j

Fig. 11.25 Mass balance open system

39 However, in open systems the mass within the system may change.
192 11 First Law of Thermodynamics

Energy supplied to the system

Heat, i.e. thermal energy Q t , can pass the system boundary regardless of whether
the system is open or closed. This can happen due to an insufficient insulation or
even due to an intended process. It is already known, that heat is initiated by a tem-
perature potential. Furthermore, according to the second law of thermodynamics, see
Sect. 14.2, heat is always transferred from a higher to a lower temperature. Conse-
quently, Q t in the following approach comprises the total heat energy transferred
due to temperature differences: This can occur both at the system boundary, i.e. for
example at the housing, and by conductive heat transfer within the fluid at inlet and
outlet. In addition to thermal energy, technical work Wt,t may also pass the system
boundary. Both, heat Q t as well as technical work Wt,t , can be positive or nega-
tive.40 In case heat or technical work are negative, energy is released by the system.
Energy is supplied to the system when they are positive.
Now the energy transferred into the system at the inlet is examined, i.e.
• Due to the velocity of the mass entering the system, kinetic energy is carried into
the system. It follows
1
m 1 c12 (11.138)
2
• Let us assume that the inlet has a vertical position of z 1 in the gravitational field.
Consequently, the mass flow has potential energy of

m 1 gz 1 (11.139)

• Furthermore, the fluid contains internal energy, see Sect. 2.4.2, i.e.

m 1 u 1 (11.140)

• At the inlet, the piece of mass m 1 is faced with a local pressure p1 at the system
boundary. In order to transfer the mass into the system, a force is required to
overcome the local pressure. This force must move the piece of mass until the
mass is completely part of the system. The corresponding work has already been
introduced in Sect. 9.2.8 and is called shifting work:

m 1 p1 v1 (11.141)

This results in the total energy supplied to the system as follows, cf. also Fig. 11.26:
 
1 2
E in = E 1 = Q t + Wt,t + m 1 c + gz 1 + u 1 + p1 v1 . (11.142)
2 1

40 Sure, both can also disappear.


11.3 Open Systems 193

Fig. 11.26 Energy crossing


the system boundary of an
open system

Specific enthalpy as new state value

In Chap. 10 the specific enthalpy as a new state value has already been introduced,
i.e.
h = u + pv. (11.143)

Obviously, the specific enthalpy h is a state value, since it is purely composed of state
values, i.e. no information regarding the path a system takes is required. Within this
chapter, i.e. for open systems, the specific enthalpy can be explained descriptively, as
it includes the specific internal energy u and the specific shifting41 work pv needed to
bring a piece of mass into a system against the local pressure p. All state values have
in common, that they characterise the internal state of a fluid42 : In an open system,
the state value specific enthalpy has proven to be useful and comparatively easy to
access. Nevertheless, the specific enthalpy represents a state value that characterises
the state of a fluid and is independent of the type of system. Consequently, the specific
enthalpy is also present in closed systems, although there is no shifting work as in
open systems.
Since specific enthalpy is a state value, a total differential need to exist, cf. Theorem
10.1. The total differential for the specific enthalpy obeys




∂h ∂h ∂h
dh = du + dv + d p. (11.144)
∂u p,v ∂v p,u ∂p u,v

41 See Sect. 9.2.8.


42 However, some state values have a vivid physical meaning, others are regarded as more or less
artificial.
194 11 First Law of Thermodynamics

Thus, the total differential reads as

dh = du + p dv + v d p. (11.145)

For ideal gases, however, it follows from Eq. 11.143 by applying the thermal equation
of state
h = u + RT. (11.146)

The total energy supplied to the system is therefore given by Eq. 11.142, i.e.
 
1 2
E in = E 1 = Q t + Wt,t + m 1 c + gz 1 + h 1 (11.147)
2 1

Energy released by the system

According to Fig. 11.26 the entire energy leaving the system follows
 
1 2
E out = E 2 = m 2 c + gz 2 + h 2 (11.148)
2 2

According to the supplied energy as given by Eq. 11.147, the energy leaving the
system is composed out of kinetic, potential, internal and shifting work. As already
mentioned, internal energy as well as shifting work are summarised as enthalpy. Nev-
ertheless, the process values heat and work may leave the system as well. However,
this case is accounted for by a negative sign for Q t or Wt,t in Eq. 11.147.

Conclusion

Now that the energy content of an open system at two different points in time and the
energy input and output across the system boundary have been clarified, Eq. 11.132
can be applied. Due to the energy conservation principle, the temporal change of a
system’s energy content depends on released and supplied energy, see Fig. 11.2743 :
 2
  2

cSys,t cSys,0
m Sys,t u Sys,t + + gz Sys,t − m Sys,0 u Sys,0 + + gz Sys,0 =
2 2



c12 c2
m 1 h1 + + gz 1 − m 2 h 2 + 2 + gz 2 + Q t + Wt,t .
2 2
(11.149)

43Again, energy is represented by bubbles, so the principle of conservation of energy is actually


reduced to counting bubbles.
11.3 Open Systems 195

Fig. 11.27 First law of thermodynamics for an open system

Equation 11.149 gives the mathematical correlation of what is illustrated in Fig. 11.27:
The change of energy, represented by bubbles, inside the system within a period of
time of t is balanced by the flow of energy, i.e. bubbles, across the system boundary.
The change of energy within the system is given in the first line of Eq. 11.149, the
flows across the boundary in the second line of that equation.

11.3.2 Non-steady State Flows

In this case, the inflows and outflows are variable over a period of time. Therefore,
the balances are as follows:

Mass balance

t t
m Sys,t − m Sys,0 = ṁ 1 (t) dt − ṁ 2 (t) dt. (11.150)
0 0

The change in mass inside the system over a period of time correlates with the mass
flows across the system boundary. An inflow increases the mass inside, an outflow
decreases the content of the system.

Energy balance

The thermal energy flow and the mechanical power can also depend on time, so that
the thermal energy yields
196 11 First Law of Thermodynamics

t
Q t = Q̇ dt (11.151)
0

respectively the technical work is

t
Wt,t = Pt dt. (11.152)
0

The convective energy fluxes with incoming and outgoing mass may also be time-
invariant, i.e.

t
t

c12 c22
E convective = h1 + + gz 1 ṁ 1 dt − h2 + + gz 2 ṁ 2 dt.
2 2
0 0
(11.153)
According to Eq. 11.149 the energy balance can be adjusted for time-invariant fluxes:
 2
  2

cSys,t cSys,0
m Sys,t u Sys,t + + gz Sys,t − m Sys,0 u Sys,0 + + gz Sys,0 =
2 2
t
t
t t
c12 c2
h1 + + gz 1 ṁ 1 dt − h 2 + 2 + gz 2 ṁ 2 dt + Q̇ dt + Pt dt.
2 2
0 0 0 0
(11.154)

Problem 11.6 In Sect. 9.2.9 an equivalent model has been introduced, see Fig. 9.15,
to derive the technical work. An open system with a constant mass flux is treated by
means of a closed system: A gas flow enters with an inlet pressure of pE and shall
leave the system with an outlet pressure of pA < pE . In the equivalent model, the
cylinder is first filled isobarically with the gas. Then the gas expands in the closed
cylinder. Finally, the gas is released isobarically to reach the desired state (A), see
Fig. 11.28. Please derive the energy balance for the entire change of state, i.e. (E) →
(A)!

Solution

The continuous flow of an open system is modelled by a so-called batch process, i.e.
a discontinuous process. Its changes of state are as follows:
(1) Isobaric filling of the reference volume: (E) → (1)
11.3 Open Systems 197

The system can be treated as an open system, since mass passes the system
boundary. Hence, Eq. 11.149 can be applied. According to this problem it reads
as



c2 c2
m 1 u 1 + 1 + gz 1 −0 = m E h E + E + gz E −0 + Q E1 + WE1 .
2 2
     
System at t Inflow
(11.155)
Example 7.6 has shown that for an isobaric process, the change of state has to
run so slowly,44 that no dissipation/friction occurs. Furthermore, the potential
energy of the gas needs to be neglected. Thus, it is

m 1 (u 1 + 0) = m E (h E + 0) + Q E1 + WE1 . (11.156)

Since the system is initially empty, the following applies to the mass balance

m 1 = m E + m E = m E . (11.157)

m E denotes the mass that enters the system. Hence, the energy balance sim-
plifies to
U1 = HE + Q E1 + WE1 . (11.158)

The partial energy equation for the work is

WE1 = WE1,V + WE1,mech + E1 (11.159)

with the mechanical work45


WE1,mech = 0 (11.160)

Fig. 11.28 Sketch to Problem 11.6, see also Fig. 9.15

44 Thus, the kinetic energy of the inflow is ignored.


45 Potential as well as kinetic energy of the gas are neglected, see Example 7.6.
198 11 First Law of Thermodynamics

and the volume work for an isobaric change of state46

WE1,V = − p1 V1 . (11.161)

Hence, the work can be summarised as47

WE1 = − p1 V1 . (11.162)

Combining these equations brings

U1 = HE + Q E1 − p1 V1 (11.163)

respectively
U1 + p1 V1 = HE + Q E1 . (11.164)
  
=H1

In other words
H1 = HE + Q E1 (11.165)

(2) Expansion of the enclosed mass: (1) → (2)

The system can be treated as a closed system, since the mass inside the system
is constant.
• The exchanged work can be calculated by applying the partial energy equation

2
W12 = − p dV + E pot,12 + 12 (11.166)
1

In this change of state, volume change work, mechanical work and also dissi-
pation occur. Section 9.2.9 is thus extended.
• The first law of thermodynamics, i.e. for a closed system, yields

Q 12 + W12 = U2 − U1 + E pot,12 (11.167)

Combining both equations brings

2
Q 12 − p dV + 12 = U2 − U1 (11.168)
1

46 The volume change lifts the piston. Thus, work is released to increase the piston’s potential
energy.
47 Mind, that there is no dissipation since the state of state needs to run very slowly, see Sect. 9.2.5!.
11.3 Open Systems 199

(3) Isobaric releasing of the entire mass: (2) → (A)

The system can be treated as an open system, since mass passes the system
boundary. Hence, Eq. 11.149 can be applied. According to this problem it reads
as



c2 c2
0 − m 2 u 2 + 2 + gz 2 = 0 − m A h A + A + gz A +Q 2A + W2A .
2 2
     
System at t=0 Outflow
(11.169)
Example 7.6 has shown that for an isobaric process, the change of state has to
run so slowly,48 that no dissipation/friction occurs. Furthermore, the potential
energy of the gas needs to be neglected. Thus, it is

− m 2 (u 2 + 0) = −m A (h A + 0) + Q 2A + W2A . (11.170)

Since the system finally is empty, the mass balance obeys

m A = m 2 − m A → m 2 = m A . (11.171)

m A denotes the mass that leaves the system. Hence, the energy balance sim-
plifies to
− U2 = −HA + Q 2A + W2A . (11.172)

The partial energy equation for the work is

W2A = W2A,V + W2A,mech + 2A (11.173)

with the mechanical work49


W2A,mech = 0 (11.174)

and the volume work for an isobaric change of state50

W2A,V = p2 V2 . (11.175)

Hence, the work can be summarised as51

W2A = p2 V2 . (11.176)

Combining these equations brings

48 Thus, the kinetic energy of the inflow is ignored.


49 Potential as well as kinetic energy of the gas are neglected, see Example 7.6.
50 The potential energy of the piston is reduced and supplied to the gas.
51 Mind, that there is no dissipation since the state of state needs to run very slowly, see Sect. 9.2.5.
200 11 First Law of Thermodynamics

− U2 = −HA + Q 2A + p2 V2 (11.177)

respectively
−U − p V = −HA + Q 2A . (11.178)
 2  2 2
=−H2

In other words
HA = H2 + Q 2A (11.179)

The summation of the energy balances of all sub-steps, i.e. Eqs. 11.165, 11.168
and 11.179, leads to the overall energy balance:

2
Q EA + EA − p dV = H1 − HE + U2 − U1 + HA − H2 (11.180)
1

with
Q EA = Q E1 + Q 12 + Q 2A (11.181)

and
EA = 12 . (11.182)

Substitution of U1 = H1 − p1 V1 and U2 = H2 − p2 V2 reads as

2
Q EA + EA − p dV = H1 − HE + H2 − p2 V2 − H1 + p1 V1 + HA − H2 .
1
(11.183)
Rearranging leads to

2
Q EA + EA − p dV = HA − HE − ( p2 V2 − p1 V1 ) . (11.184)
1

Applying Eq. 9.112 yields the energy balance of the entire process (E) → (A)

2
Q EA + V d p + EA = HA − HE . (11.185)
1

Since the first and last steps are isobaric, the following still applies
11.3 Open Systems 201

2 A
V dp = V d p. (11.186)
1 E

Thus, the first law of thermodynamics finally obeys

A
Q EA + V d p + EA = HA − HE (11.187)
E

Problem 11.7 A hot air flow ṁ (u 1 = 250 kg kJ


, ρ1 = 2.0 mkg3 , p1 = 2 bar) flows
through a throttle valve into a large tank. The velocity in the inlet pipe is c = 10 m s−1 .
The cross section of the inlet pipe is A = 1 × 10−2 m2 and its height above ref-
erence level z 0 is 10 m. To ensure homogeneity in the tank, an impeller with a
power consumption of 10 kW is mounted. The tank is at rest, i.e. its velocity is
zero, and the reference level, i.e. z 0 = 0, is at the tank’s centre of gravity, cf.
Fig. 11.29. Calculate the heat exchanged with the environment for a time interval
of t = 10 s so that the specific internal energy of the tank contents remains con-
stant (u Sys = 214.2 kg
kJ
= const.).

Solution

Obviously, the problem at hand is an open system, as mass enters the system. How-
ever, this particular problem cannot occur in the steady state because no mass leaves
the system. Consequently, the mass inside increases. Figure 11.29 shows a sketch of
the problem. Let us now apply Eq. 11.149 under the assumption that m 2 = 0, since
no mass flow leaves the system:

Fig. 11.29 Sketch to


Problem 11.7
202 11 First Law of Thermodynamics
 2
  2

cSys,t cSys,0
m Sys,t u Sys,t + + gz Sys,t − m Sys,0 u Sys,0 + + gz Sys,0 =
2 2


c12
m 1 h 1 + + gz 1 + Q + Wt .
2
(11.188)
Note that m Sys,t is the mass of the system after t = 10 s and m Sys,0 is the initial
mass which is unknown in this case. The centre of gravity of the system is at rest and
does not vary vertically because the air, as an ideal gas, always occupies the entire
volume. Thus, the first law of thermodynamics simplifies to


c12
m Sys,t u Sys,t − m Sys,0 u Sys,0 = m 1 h1 + + gz 1 + Q + Wt . (11.189)
2

According to Eq. 11.150 the mass balance follows

t
m Sys,t − m Sys,0 = ṁ 1 (t) dt = ṁ 1 t = m 1 . (11.190)
0

Since the specific internal energy shall be constant, i.e. u Sys,t = u Sys,0 = u Sys =
const., the first law of thermodynamics obeys


  c12
u Sys m Sys,t − m Sys,0 = m 1 h1 + + gz 1 + Q + Wt . (11.191)
   2
=m 1

The following terms can be calculated:


• Technical work Wt
Wt = Pt t = 100 kJ (11.192)

• Mass m 1
m 1 = ρ1 V̇1 t = ρ1 c1 A t = 2 kg (11.193)

• Specific outer energies


c12 J
+ gz = 148 (11.194)
2 kg

• Specific enthalpy h 1

p1 kJ
h 1 = u 1 + p1 v1 = u 1 + = 350 (11.195)
ρ1 kg

Thus, the heat obeys


11.3 Open Systems 203



c2
Q = u Sys m 1 − m 1 h 1 + 1 + gz 1 − Wt = −371.9 J (11.196)
2

11.3.3 Steady State Flows

In this section open systems in steady state are investigated, i.e. systems that do not
vary with time. The following premises apply to stationary, open systems:
1. The mass of a system needs to be constant, i.e. m Sys,t = m Sys,0 . With respect to
Eq. 11.137 it implies that
 
m in,i = m out, j (11.197)
i j

respectively  
ṁ in,i = ṁ out, j . (11.198)
i j

If the incoming and outgoing masses were different, the mass in the system would
either increase or decrease.
2. All fluxes across the system boundary are time-independent, i.e. mass fluxes as
well as energy fluxes.
3. All state values within the system need to be temporally constant, i.e.
• Specific respectively extensive52 state values

z = f (t) → Z = f (t) (11.199)

• Intensive state values


p = f (t), T = f (t) (11.200)

• Outer state values


z Sys = const., cSys = const. (11.201)

However, state values can vary locally inside an open system. Pressure, as inten-
sive state value, for instance can be different in an open system at inlet and outlet.
Consequently, the pressure develops from inlet to outlet. Pressure sensors, located
at different positions inside the system, would prove that. In case a steady state
occurs, though the state values are locally different, there is no change with time,
i.e. at any time the local sensor measures constant values. Anyhow, the system is
supposed to be in local equilibrium, see Sect. 4.4. Thus, the energy content of the
system is constant with time, i.e.

52 Note that the mass of the system is constant.


204 11 First Law of Thermodynamics

E Sys,t = E Sys,0 (11.202)

According to Eq. 11.132, an open system in steady state obeys53


 
E Sys,t − E Sys,0 = 0 = E in,i − E out, j (11.203)
i j

If the open system consists of an inlet and an outlet, the energy balance can be derived
from Eq. 11.154 and reads as follows for the premises made



c2 c2
0 = h 1 + 1 + gz 1 ṁ 1 t − h 2 + 2 + gz 2 ṁ 2 t + Q̇ t + Pt t
2 2
(11.204)
with
ṁ 1 = ṁ 2 = ṁ. (11.205)

Hence, it is



c12 c22
0 = ṁ h 1 + + gz 1 − ṁ h 2 + + gz 2 + Q̇ + Pt . (11.206)
2 2

Finally, the first law of thermodynamics for open systems with one inlet/outlet in
steady state obeys
 
1 2 
Q̇ + Pt = ṁ (h 2 − h 1 ) + c2 − c12 + g (z 2 − z 1 ) (11.207)
2

The specific notation results from the division by the constant mass flux ṁ:

1 2 
q12 + wt,12 = (h 2 − h 1 ) + c − c12 + g (z 2 − z 1 ) (11.208)
2 2

Note that Q̇ = ṁq12 and Pt = ṁwt,12 .

Open systems in steady state with multiple inlets/outlets

The focus is now on open systems in steady state with multiple inlets/outlets. Apply-
ing Eq. 11.132 leads to:
 
E Sys,t − E Sys,0 = 0 = E in,i − E out, j (11.209)
i j

53 This is due to the energetic state of a system does not vary by time in steady state.
11.3 Open Systems 205

Fig. 11.30 Open systems


with multiple inlets/outlets

respectively  
E in,i = E out, j . (11.210)
i j

The derivation with respect to time is


 
Ė in,i = Ė out, j (11.211)
i j

In other words:

Energy flux in = Energy flux out

This principle can be easily adapted to any open system in steady state. For the
example illustrated in Fig. 11.30 it reads as54



c12 c22
Q̇ 12 + Pt12 + ṁ 1 h1 + + gz 1 + ṁ 2 h 2 + + gz 2 =
2 2

(11.212)
c2
ṁ 3 h 3 + 3 + gz 3 .
2

In combination with the mass balance55

Mass flux in = Mass flux out

it is for the shown example, see Fig. 11.30:

ṁ 1 + ṁ 2 = ṁ 3 . (11.213)

Hence, the first law of thermodynamics obeys

54 Note that arrows pointing into the system are energy fluxes in, while arrows pointing out of the
system represent energy fluxes out. Anyhow, a mass flux carries enthalpy, potential and kinetic
energy.
55 In steady state.
206 11 First Law of Thermodynamics
 
c32 − c12
Q̇ 12 + Pt12 = ṁ 1 (h 3 − h 1 ) + + g (z 3 − z 1 ) +
2
  (11.214)
c32 − c22
ṁ 2 (h 3 − h 2 ) + + g (z 3 − z 2 ) .
2

This notation, i.e. differences of state values on the right hand side, is advantageous,
see Chap. 12. In this case, the caloric equations of state can be easily applied.56

11.3.4 Partial Energy Equation

Similar to Sect. 11.2.3, where the partial energy equation for the process value work
for closed systems was derived, the technical work wt , which is characteristic for
open systems, see Eq. 11.208, is now further investigated. A partial energy equation
is not a complete energy balance, but it illustrates how work is converted by a system.
Note that caloric state values such as specific internal energy/enthalpy/entropy are
not covered by the partial energy equation, but occur in the first respectively second
law of thermodynamics. Anyhow, for an open system in steady state the first law of
thermodynamics obeys
 
c2 − c12
Q̇ 12 + Pt12 = ṁ h 2 − h 1 + 2 + g · (z 2 − z 1 ) (11.215)
2

respectively by dividing by the mass flux

c22 − c12
q12 + wt12 = h 2 − h 1 + + g · (z 2 − z 1 ) . (11.216)
2
Mind, that q12 as well as wt12 only take into account energies that cross the system
boundary. It is therefore important to clearly define the system boundaries. Never-
theless, technical work is only relevant for the first law of thermodynamics, when
mechanical or electrical work is obviously transferred: This can be the case, for
example, when a mechanical shaft crosses the system boundary, see Fig. 11.31. Such
a shaft can supply the system with energy,57 or the flow inside the systems drives the
shaft.58 Furthermore, electrical energy can be supplied to a system, so that technical
work is positive. If there are no such electrical wires or mechanical shafts, e.g. a
section of heating pipe without the water pump, then Pt = ṁwt12 = 0.

56 Caloric equations of state indicate how the change in a caloric state value can be calculated. Note
that caloric state values, such as specific enthalpy or specific internal energy, cannot be measured
but must be calculated.
57 So the technical work is positive.
58 So the technical work is negative.
11.3 Open Systems 207

Fig. 11.31 Technical work

In order to derive the partial energy equation for the specific technical work, a mass
particle dm is regarded within an open system. Let this piece of mass be connected
with a moving coordinate system, see Fig. 11.31. However, this mass particle can be
treated as a closed system, since its mass shall be constant on its way from inlet (1)
to outlet (2).
Thus, the first law of thermodynamics for closed systems, cf. Eq. 11.20, can be
applied for the particle dm. In differential notation it obviously reads as:

δq + δw = du + dea . (11.217)

Applying the partial energy equation for the specific work of a closed system, accord-
ing to Sect. 11.2.3, leads to

δw = δwV + δwmech + δψ = − p dv + dea + δψ. (11.218)

The particle, treated as a closed system, is obviously subject to volume work, since it
may be compressed or expand while its mass remains constant. Furthermore, work is
required for lifting and acceleration its centre of gravity. Note that in Eq. 11.218 the
changes in outer energies related to the centre of gravity require mechanical work.59
Finally, energy may be dissipated due to internal friction.
The specific enthalpy of the mass particle follows, see Eq. 11.145:

dh = du + p dv + v d p. (11.219)

Rearranging for du leads to:

du = dh − p dv − v d p. (11.220)

59 According to Eq. 11.22 lifting or accelerating the centre of gravity requires mechanical work.
208 11 First Law of Thermodynamics

Hence, it is
δq + δψ = dh − v d p. (11.221)

Following the particle from (1) to (2) means integrating Eq. 11.221, so that

2
q12 + ψ12 = h 2 − h 1 − v d p. (11.222)
1

Solving for the specific heat leads to:

2
q12 = h 2 − h 1 − v d p − ψ12 . (11.223)
1

Replacing the specific heat in Eq. 11.216 for an open system, i.e. joining the two
coordinate systems,60 leads to the desired equation for the technical work:

2
c22 − c12
wt12 = v d p +ψ12 + + g (z 2 − z 1 ) (11.224)
1
 2  
   ea12
y12

Example 11.3 Fortunately, Eq. 11.224 can be visualised: think of a high-pressure


pump that is to be operated with water from a lower-lying lake. At the inlet the
velocity is zero, since the lake is at rest. At the outlet, which is at a higher vertical
position than the inlet, the fluid shall be accelerated to a velocity c2 . Furthermore, its
pressure shall be increased as well, see Fig. 11.32. The question now is what power
the pump must have. Anyhow, Eq. 11.224 explains how the required technical work
is composed:
• Specific work to increase the pressure, i.e. specific pressure work y12

2
y12 = v d p. (11.225)
1

Pressure work only occurs in open systems.61

60 The moving coordinate system represents the single particle, while the fixed coordinate system
represents the integral energy balance as given by Eq. 11.216.
61 In contrast, volume work only occurs in closed systems. However, both volume and pressure

work correspond to the physical definition of work, i.e. force times distance.
11.3 Open Systems 209

Fig. 11.32 What does technical work mean?

Fig. 11.33 Specific pressure work y12

• Due to dissipation ψ12 inside the system the required specific work is increased.62
c2 −c2
• Specific work to increase the kinetic energy of the fluid 2 2 1 . The velocity at the
inlet is supposed to be zero, cf. Fig. 11.32. Even if the velocity tends towards zero,
mass flow can still occur because the cross-section of the inlet is supposed to be
large.
• Specific work to increase the vertical position of the fluid g (z 2 − z 1 ).
The next step is to investigate how to calculate the specific pressure work y12 .
Therefore, the states (1) and (2) are first shown in a p, v-diagram, see Fig. 11.33.
2
The integral v d p is represented by the projected area to the p-axis as indicated
1
by Fig. 11.33. The p, v-diagram can thus be applied not only to illustrate the specific
volume work wV in a closed system, but also the specific pressure work y12 for open
systems. Similar to closed systems, the path (1) to (2) must also be known for open
systems in order to calculate the specific pressure work. The specific pressure work
is therefore a process variable, just like the specific volume work for closed systems.
Theorem 11.3 The pressure work is part of the technical work in an open system.
It describes the work that must be supplied in order to increase the pressure of the
fluid—or the work that is released by a fluid when the pressure drops.

62 This is well known from daily experience: Systems with friction require more effort than fric-
tionless systems.
210 11 First Law of Thermodynamics

In Sect. 9.2.9 the pressure work has already been derived on an alternate way.
This alternative approach was based on a cylinder/piston system with inlet and out-
let valves, which could be described by means of a closed system. However, both
approaches lead to the same correlation for the pressure work.

Why do we need the partial energy equation?

The partial energy equation

2
c22 − c12
wt12 = v d p + ψ12 + + g (z 2 − z 1 ) (11.226)
2
1

should be applied when pressure change in open systems and/or dissipation are
required. This can be illustrated with the following example: Imagine a horizontal
section of pipe through which a fluid flows in a steady state. Let us further assume
that the changes in kinetic energies can be neglected. Under these conditions, the first
law of thermodynamics for open, stationary systems satisfies the following equation

Q̇ 12 + Pt12 = ṁ (h 2 − h 1 ) . (11.227)

As can be seen, the first law of thermodynamics does not include any information
regarding the pressure as well as the dissipation. Obviously, the partial energy equa-
tion contains such information and, with the assumptions made, simplifies to

2
wt12 = v d p + ψ12 . (11.228)
1

However, since the tube section is a work-insulated system, no work crosses the
system boundary,63 so that

2
wt12 = 0 = v d p + ψ12 . (11.229)
1

This leads to
2
v d p = −ψ12 ≤ 0. (11.230)
1

63 Note, as documented in Fig. 11.31, technical work crosses the system boundary only when there
is a mechanical shaft, electrical wires or similar to transport energy in form of work into or out of
the system. This is not the case—the presented example purely is a passive tube section.
11.3 Open Systems 211

The right side of this equation is negative because the specific dissipation is always
ψ12 ≥ 0. It is obvious that specific dissipation causes a pressure drop under the given
conditions. If there is no dissipation, the flow is isobaric. If dissipation occurs, a
pressure drop is measured because

2
v d p < 0. (11.231)
1

This is only possible if


d p < 0. (11.232)

The integration from (1) to (2) reads as

p2 − p1 < 0 (11.233)

respectively
p2 < p1 . (11.234)
Chapter 12
Caloric Equations of State

In the previous chapters, the law of conservation of energy, which enables ther-
modynamic systems to be energetically evaluated, has been discussed in detail. A
distinction has been made between closed and open systems. However, in addition
to the easily measurable thermal state values, such as pressure p or temperature T , a
new category of state values has been introduced: These state values, e.g. the specific
internal energy u and the specific enthalpy h, cannot be determined by a sensor and
must therefore be calculated. They are denoted caloric state values and occur, for
example, in the first law of thermodynamics:
• Closed systems
q12 + w12 = u 2 − u 1 + Δea12 (12.1)

• Open systems
q12 + wt12 = h 2 − h 1 + Δea12 (12.2)

Caloric state values have in common that they all contain an energy dimension.1
In this chapter, the caloric state values are further investigated and equations of
state are derived to calculate and predict them. In addition, a new state value, the
specific entropy s, is introduced. This state value quantifies the constraint on energy
conversion, see Chap. 15. Anyway, in Sect. 5.1, Gibbs’ phase rule has been discussed,
which has proved that two independent specific state values are required to uniquely
define a state of an ideal gas. Accordingly, the following mathematical correlations
can be stated2 :
u = u (v, T ) (12.3)

1 Such as [u] = 1 kJ
kg and [h] = 1 kJ
kg .
2If two independent state values unambiguously define the thermodynamic state, all other state
values must also unambiguously belong to this state and be determinable.

© Springer Nature Switzerland AG 2022 213


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_12
214 12 Caloric Equations of State

h = h ( p, T ) (12.4)

s = s ( p, T ) . (12.5)

However, the choice of which dependencies are specified in Eqs. 12.3 to 12.5 is
arbitrary. It is also possible to apply different functions, e.g. u = u ( p, T ), but for
the thermodynamic derivations to come, e.g. the introduction of the specific heat
capacities, it makes sense to proceed with the given dependencies. Subsequently, it
is possible to formulate other correlations, e.g.

h = h (s, v) . (12.6)

12.1 Specific Internal Energy u and Specific Enthalpy h for


Ideal Gases

The caloric equations of state for ideal gases are indeed rather simple and are briefly
summarised here. For those readers interested in a more detailed thermodynamic
derivation, please refer to Sect. 12.3. The total differentials3 for the specific inter-
nal energy and the specific enthalpy result from Eqs. 12.3 and 12.4. According to
Chap. 10, these total differentials exist, since u and h are state values, see Theo-
rem 10.1:
• Specific internal energy u
   
∂u ∂u
du = · dT + · dv (12.7)
∂T v ∂v T

The first partial derivative is defined as follows


 
∂u
= cv (v, T ) (12.8)
∂T v

cv is called specific heat capacity.4 Since u is a function of two independent state


values, i.e. in this case v and T , cv also depends on these values.
• Specific enthalpy h
   
∂h ∂h
dh = · dT + · dp (12.9)
∂T p ∂p T

3 The indices at the brackets indicate that this variable is kept constant.
4 At constant specific volume v.
12.1 Specific Internal Energy u and Specific Enthalpy h for Ideal Gases 215

The first partial derivative is defined as follows


 
∂h
= c p ( p, T ) (12.10)
∂T p

c p is called specific heat capacity.5 Since h is a function of two independent state


values, i.e. in this case p and T , c p also depends on these values.
According to [21], experiments, e.g. by Joule, cf. Example 12.1, have shown that for
ideal gases the specific internal energy and the specific enthalpy depend only on the
temperature, i.e.6 :  
∂u
=0 (12.11)
∂v T

and  
∂h
= 0. (12.12)
∂p T

Since u and h are merely functions of the temperature T , so are the specific heat
capacities cv and c p . This results in the following caloric equations of state for ideal
gases:
du = cv (T ) dT (12.13)

respectively
dh = c p (T ) dT (12.14)

The specific heat capacities cv and c p of the fluid and the handling of the caloric
equations of state are explained in more detail in Sect. 12.4.

In summary, for ideal gases the caloric equations of state, i.e. Eqs. 12.13 and 12.14,
as well as the thermal equation of state, i.e.

pv = RT, (12.15)

have been validated by experiments, see Chap. 6 and Example 12.1.


Example 12.1 In this example the Joule expansion, cf. [22], is introduced, see
Fig. 12.1. Two vessels containing an ideal gas are connected by a pipe which is
closed by a valve in state (1). The pressure in vessel B shall be lower than the pres-
sure in A. At the beginning, however, the two vessels are in thermal equilibrium with
the environment, i.e. each gas has the same temperature T . The entire system is now
adiabatically isolated. By opening the valve, a mechanical balancing process begins.
In state (2) the two vessels are in thermodynamic equilibrium, i.e. a new pressure p2

5 At constant pressure p.
6 The thermodynamic proof is given in Sect. 12.3.2.
216 12 Caloric Equations of State

Fig. 12.1 Joule expansion

is reached in both vessels. Joule found that the change of state is isothermal, i.e. the
temperature remains constant for any chosen pressure difference between A and B.
According to the first law of thermodynamics it is7

W12 + Q 12 = 0 = U2 − U1 = ΔUA + ΔUB = m A (u 2 − u A ) + m B (u 2 − u B )


(12.16)
respectively
u 2 = u A = u B = const. (12.17)

Obviously, pressure has no influence on the specific internal energy. If pressure had
an influence, the specific internal energy in (1) and (2) would have to be different,8
since the pressures in (1) and (2) are different, i.e.

u = f ( p) . (12.18)

The same applies to the specific volume, which changes from state (1) to state (2),
so that further
u = f (v) . (12.19)

It must therefore apply that


 
∂u
=0 (12.20)
∂v T

7 Note that the system is adiabatic, i.e. Q 12 = 0, and that no work is exchanged across the system
boundary, i.e. W12 = 0. The work to open the valve is neglected.
8 Yet this would violate Eq. 12.17.
12.1 Specific Internal Energy u and Specific Enthalpy h for Ideal Gases 217

The first law of thermodynamics and the observation that T remains constant thus
lead to9
u = f (T ) . (12.21)

For an ideal gas it further is

h = u + pv = u (T ) + RT. (12.22)

Thus, this yields


h = f (T ) . (12.23)

It must therefore apply that


 
∂h
=0 (12.24)
∂p T

However, in case the deviations from the ideal gas behaviour increase, the change
of state is no longer isothermal, i.e. the specific internal energy is a function of
temperature and pressure, see also Sect. 18.5.

12.2 Specific Entropy s as New State Value for Ideal Gases

In this section, a new state value is introduced. The first part of this section, i.e. the
mathematical derivation, is rather theoretical. In Chap. 13, however, this new state
value is thermodynamically motivated with simple examples.
At this point, the question arises as to why a new state value is essential for under-
standing energy conversion, since the first law of thermodynamics, which demon-
strates the principle of energy conservation, has already been discussed. Unfortu-
nately, the first law of thermodynamics does not explain the restrictions on the direc-
tion of energy conversion: According to the first law of thermodynamics, it would
be possible, for example, for the temperature of a hot cup of tea in a room with low
ambient temperature to rise further while the room cools down. In other words, the
energy supplied to the tea is taken from the environment, i.e. the amount of energy
would be constant and the first law of thermodynamics would be satisfied. However,
everyday experience rather shows the opposite, namely that the tea cools down. From
this point of view, it is obvious that the first law of thermodynamics on its own is not
sufficient for a comprehensive understanding. In order to understand the direction of
energy transformation, a new state value, namely entropy, is required.
Before starting with the theoretical derivation, it is important to distinguish once
again between state and process value, see Chap. 10:

9If T remains constant, the specific internal energy does not change. This is exactly what was
observed in the experiment.
218 12 Caloric Equations of State

Theorem 12.1 A state value describes the state of a system and does not take into
account the path the system follows to reach this state. In contrast to a state value,
a process value takes into account the path that the system follows from an initial
state to a new state.
Volume work in a closed system, respectively pressure work in an open system are
process values, since they only occur, when the system underlies a change of state.
However, a thermodynamic state is not associated with work or heat. Process values,
i.e. heat, work and dissipation, depend on the path the system follows. In contrast to
that, state values characterise the system path-independently. Some state values have
a comprehensible physical meaning, as they are part of our daily life, e.g. temperature
T or pressure p. Others are more difficult to interpret, e.g. specific internal energy.
Some others even appear to be artificial: Specific enthalpy was introduced with open
systems, but like all state values it also occurs in closed systems. State values describe
the fluid. It has been shown that the combination of state values leads to new state
values, see Chap. 10. When deriving the caloric equations of state in Sect. 12.3, two
new state values appear which so far have no comprehensible physical meaning:
The specific Helmholtz energy f and the specific Gibbs energy g. As for these two
state values, they receive a physical interpretation as soon as the chemical reaction
systems are introduced in Chap. 24.
Finally, a new state value is now derived, which accompanies us until the end of
this book. Experience has shown that this new state value entropy s causes difficulties
because, as an engineer, one is trying to find a physical motivation right from the
beginning.10 It is easiest to first accept the specific entropy s as a new state value. In
Chap. 13 the specific entropy is applied and its meaning becomes more accessible.
Finally, the restriction of energy conversion, e.g. in a power plant, can be motivated
by this new state value and its characteristics. By the end of Part I, most readers will
have become accustomed to this new state value.
The first law of thermodynamics for a closed system in differential, specific nota-
tion reads as
δq + δw = du + dea . (12.25)

With the partial energy equation it is possible to specify the specific work δw, i.e.

δw = − p dv + δψ + δwmech . (12.26)
  
=dea

Combining these two equations leads to

δq + δψ = du + p dv. (12.27)

10Frankly, it is even difficult to find a physical meaning for the inner energy. Even though you have
probably accepted its existence by now.
12.2 Specific Entropy s as New State Value for Ideal Gases 219

By introducing the specific enthalpy h, see Eq. 11.145:

du = dh − p dv − v d p (12.28)

the equation also reads as


δq + δψ = dh − v d p. (12.29)

Equation 12.27 can be divided11 by the temperature T , so that

δq + δψ du p dv
= + . (12.30)
T T T

Equation 12.29 can alternatively be derived with the first law of thermodynamics for
open systems, i.e.
δq + δwt = dh + dea . (12.31)

The partial energy equation for the technical work δwt is

δwt = v d p + δψ + dea . (12.32)

Combining these two equations leads to

δq + δψ = dh − v d p. (12.33)

Dividing by the temperature T (integrating factor) leads alternatively to

δq + δψ dh v dp
= − . (12.34)
T T T

With the caloric and the thermal equations of state for ideal gases

du = cv dT (12.35)

respectively
p R
= (12.36)
T v
Equation 12.30 obeys
δq + δψ dT dv
= cv +R . (12.37)
T T v

11 Also known as integrating factor.


220 12 Caloric Equations of State

The integration of the right side of Eq. 12.37 is path-independent, since

2 2
dT dv T2 v2
cv + R = cv ln + R ln . (12.38)
T v T1 v1
1 1

Obviously, the result of this integration depends only on the initial and final state.
No information is needed regarding the path from state (1) to state (2), i.e. it has the
quality of a state value. Thus, the left side of Eq. 12.30 is a state value as well. This
new state value is denoted as specific entropy s:

δq + δψ du p dv
ds = = + . (12.39)
T T T

Alternatively, by applying thermal and caloric equations of state Eq. 12.34 obeys:

δq + δψ dT dp
= cp −R . (12.40)
T T p

Accordingly, the integration of the right side of Eq. 12.40 is path-independent as


well, since
2 2
dT dp T2 p2
cp − R = c p ln − R ln . (12.41)
T p T1 p1
1 1

Obviously, the result of this integration depends only on the initial and final state.
No information is needed regarding the path from state (1) to state (2), i.e. it has the
quality of a state value. Thus, the left side of Eq. 12.34 is a state value as well. This
state value is denoted as specific entropy s:

δq + δψ dh v dp
ds = = − . (12.42)
T T T

Hence, Eqs. 12.39 and 12.42 are equal:

δq + δψ dh v dp du p dv
ds = = − = + . (12.43)
T T T T T
The following definition is made:

δq δψ
ds = + ≡ δsa + δsi (12.44)
T T
12.2 Specific Entropy s as New State Value for Ideal Gases 221

According to this equation, the state value specific entropy s is influenced by two
mechanisms12 :
1. The process value
δq
δsa = (12.45)
T

that is related to heat passing the system boundary. In extensive notation, i.e. by
multiplying with the mass of the system, it is

δQ
δSa = . (12.46)
T
This equation can be divided by dt, so that it follows

δSa 1 δQ
= . (12.47)
dt T dt
This equation represents derivatives with respect to time: A heat flux causes a
flux of entropy:

Ṡa = (12.48)
T

2. The process value


δψ
δsi = (12.49)
T

that is related to the dissipation within the system. In extensive notation, i.e. by
multiplying with the mass of the system, it is

δ
δSi = . (12.50)
T
This equation can be divided by dt, so that it follows

δSi 1 δ
= . (12.51)
dt T dt
This equation represents derivatives with respect to time: A flux of dissipated
energy causes a flux of entropy:

˙

Ṡi = (12.52)
T

12Similar to the first law of thermodynamics: The state value internal energy is influenced by the
process values work and heat.
222 12 Caloric Equations of State

However, Eqs. 12.48 and 12.52 are important, when open systems are treated in
Sect. 13.7.2.
The change of entropy can be calculated by integration, cf. Eq. 12.38:

T2 v2
Δs = s2 − s1 = cv ln + R ln (12.53)
T1 v1

Or, alternatively, according to Eq. 12.41:

T2 p2
Δs = s2 − s1 = c p ln − R ln (12.54)
T1 p1

Both equations, Eqs. 12.53 and 12.54, are equivalent and part of the caloric equations
of state for ideal gases. Anyhow, the conclusion of this theoretical derivation of the
specific entropy s is:
Theorem 12.2 The specific entropy s has been derived as new state value. It is
influenced by heat crossing the system boundary and by dissipation within the system.
In Chap. 13 this new state value is further motivated and it is shown how to apply
the specific entropy s.
Theorem 12.3 The conclusion that the specific entropy s is a state value, however,
does not depend on the type of fluid. Any state value can be determined for a system,
regardless of the fluid.
However, the formulation of the caloric equations of state depends on the type of
fluid. In this section, such caloric equations of state have been derived for an ideal
gas. The entropy of real fluids is further investigated in the following section. As
already mentioned, real fluids are dealt with in Part II.

12.3 Derivation of the Caloric Equations for Real Fluids

Note that this section is for advanced readers. It is possible to skip this section and
continue with Sect. 12.4. Specifically, correlations for the specific internal energy u,
the specific enthalpy h and the specific entropy s are thermodynamically derived in
this section.

12.3.1 Specific Internal Energy u

For the caloric state value specific internal energy

u = u(v, T ) (12.55)
12.3 Derivation of the Caloric Equations for Real Fluids 223

the total differential yields


  
∂u ∂u
du = dT + dv. (12.56)
∂T v ∂v T
  
cv

As mentioned before, the specific heat capacity cv is a function of v and T , since,


according

to Eq. 12.55, u depends on v and T . The goal is now to find a correlation
for ∂u
∂v T
for real fluids.

In Sect. 12.2 Eq. 12.43 has been derived. This can be mathematically transformed
and is called fundamental equation of thermodynamics:

du = T ds − p dv (12.57)

respectively
dh = T ds + v d p (12.58)

Any fluid obeys the fundamental equation of thermodynamics, since no restrictions


have been made so far. At this point it is important to emphasise that although the
specific enthalpy h has been motivated in an open system, h is a state value that can
be determined for any system, i.e. also in a closed system.

Combining Eqs. 12.56 and 12.57 leads to


   
∂u ∂u
dT + dv = T ds − p dv. (12.59)
∂T v ∂v T

A mathematical reformulation leads to


   
1 ∂u 1 ∂u
ds = dT + p+ dv. (12.60)
T ∂T v T ∂v T
  
cv

The specific entropy s is a state value, see Sect. 12.2. It therefore depends on two
independent state values for a single component/single phase fluid, cf. Sect. 5.1, e.g.

s = s(v, T ). (12.61)

Since a total differential exists for all state values, see Theorem 10.1, the following
applies    
∂s ∂s
ds = dT + dv. (12.62)
∂T v ∂v T
224 12 Caloric Equations of State

Comparison of the coefficients for dT and dv in Eqs. 12.60 and 12.62 leads to:
   
∂s 1 ∂u
= p+ (12.63)
∂v T T ∂v T

and    
∂s cv 1 ∂u
= = . (12.64)
∂T v T T ∂T v

By rearranging Eq. 12.63 yields


   
∂u ∂s
=T − p. (12.65)
∂v T ∂v T
∂s

To proceed, ∂v T
in Eq. 12.65 must be specified. For this purpose, the so-called
specific Helmholtz free energy f is defined as

f = u − T s. (12.66)

Obviously, the specific Helmholtz free energy f is a state value because it is com-
posed exclusively of state values. It is therefore path-independent. Anyhow, the total
differential obeys
d f = du − T ds − s dT. (12.67)

It follows with Eq. 12.57 that

d f = −s dT − p dv. (12.68)

Since the Helmholtz specific free energy f is a state value, it can be determined by
two independent state values, e.g. as:

f = f (T, v). (12.69)

According to that definition the total differential is


   
∂f ∂f
df = dT + dv. (12.70)
∂T v ∂v T

The comparison of the coefficients of Eqs. 12.68 and 12.70 leads to:
 
∂f
= −s (12.71)
∂T v

and  
∂f
= − p. (12.72)
∂v T
12.3 Derivation of the Caloric Equations for Real Fluids 225

The second derivatives of Eqs. 12.71 and 12.72 are as follows


⎛  ∂f  ⎞
∂ ∂T  
⎝ v⎠ ∂2 f ∂s
= =− (12.73)
∂v ∂v ∂ T ∂v T
T

and ⎛ ∂f  ⎞
∂ ∂v  
⎝ T⎠ ∂2 f ∂p
= =− . (12.74)
∂T ∂ T ∂v ∂T v
v

According to Schwarz’s Theorem 10.3, see Sect. 10.2, it follows, that Eqs. 12.73 and
12.74 must be identical,
∂s
otherwise it would not be a state function. This leads to the
required derivative ∂v :
T    
∂s ∂p
= . (12.75)
∂v T ∂T v

A combination of Eqs. 12.65 and 12.75 yields


   
∂u ∂p
=T − p. (12.76)
∂v T ∂T v

In other words, the caloric Eq. 12.56 for the specific internal energy u is:
 
∂p
du = cv (v, T ) · dT + T − p · dv (12.77)
∂T v

Furthermore, for the specific entropy s it follows from the combination of Eqs. 12.62,
12.64 and 12.75  
cv ∂p
ds = · dT + · dv (12.78)
T ∂T v

12.3.2 Specific Enthalpy h

For the caloric state value specific enthalpy

h = h( p, T ) (12.79)

the total differential obeys


   
∂h ∂h
dh = dT + d p. (12.80)
∂T p ∂p T
  
cp
226 12 Caloric Equations of State

As mentioned before, the specific heat capacity c p is a function of p and T , since


according
  to Eq. 12.79 h depends on p and T . The goal is now to find a correlation
∂h
∂p
.
T

Combining Eq. 12.80 and the fundamental equation of thermodynamics Eq. 12.58
leads to    
∂h ∂h
dT + d p = T ds + v d p. (12.81)
∂T p ∂p T

A mathematical reformulation leads to


   
1 ∂h 1 ∂h
ds = dT + − v d p. (12.82)
T ∂T p T ∂p T

The entropy as new state value also depends on two independent state values for
a single component/single phase fluid, cf. Sect. 5.1, e.g.

s = s( p, T ). (12.83)

Since s is a state value, its total differential exists:


   
∂s ∂s
ds = dT + d p. (12.84)
∂T p ∂p T

Comparison of the coefficients for dT and d p in Eqs. 12.82 and 12.84 leads to:
   
∂s 1 ∂h
= −v (12.85)
∂p T T ∂p T

and    
∂s cp 1 ∂h
= = . (12.86)
∂T p T T ∂T p

By rearranging Eq. 12.85 yields


   
∂h ∂s
=T + v. (12.87)
∂p T ∂p T
 
To proceed, ∂∂sp in Eq. 12.87 needs to be specified. For this purpose, the so-
T
called Gibbs free energy13 g is defined as

g = h − T s. (12.88)

13For the Gibbs free energy there is a physical motivation for chemical reactive systems, e.g. fuel
cells or Lithium Ion batteries. This is treated in Part III of this book, see Sect. 24.3.
12.3 Derivation of the Caloric Equations for Real Fluids 227

Obviously, the specific Gibbs free energy g is a state value because it is composed
exclusively of state values. It is therefore path-independent. Anyhow, the total dif-
ferential obeys
dg = dh − T ds − s dT. (12.89)

It follows with Eq. 12.58 that

dg = v d p − s dT. (12.90)

Since the specific Gibbs free energy g is a state value, it can be determined by two
independent state values, e.g. as:

g = g(T, p). (12.91)

According to that definition the total differential is


   
∂g ∂g
dg = dT + d p. (12.92)
∂T p ∂p T

The comparison of the coefficients of Eqs. 12.90 and 12.92 leads to:
 
∂g
= −s (12.93)
∂T p

and  
∂g
= v. (12.94)
∂p T

The second derivatives of Eqs. 12.93 and 12.94 are as follows


⎛   ⎞
∂ ∂g  
⎜ ∂T p ⎟ ∂2g ∂s
⎝ ⎠ = =− (12.95)
∂p ∂p ∂T ∂p T
T

and ⎛  ∂g  ⎞
∂ ∂p  
⎝ T⎠ ∂2g ∂v
= = . (12.96)
∂T ∂T ∂p ∂T p
p

According to Schwarz’s Theorem 10.3, see Sect. 10.2, it follows, that Eqs. 12.95 and
12.96 must be identical,
 otherwise it would not be a state function. This leads to the
required derivative ∂∂sp :
T
   
∂s ∂v
− = . (12.97)
∂p T ∂T p
228 12 Caloric Equations of State

A combination of Eqs. 12.87 and 12.97 yields


   
∂h ∂v
= −T + v. (12.98)
∂p T ∂T p

In other words, the caloric Eq. 12.80 for the specific enthalpy h is:
   
∂v
dh = c p ( p, T ) · dT + −T + v · dp (12.99)
∂T p

Furthermore, for the specific entropy s it follows from the combination of Eqs. 12.84,
12.86 and 12.97  
cp ∂v
ds = · dT − · dp (12.100)
T ∂T p

Consequences for ideal gases

So far, no constraints regarding the fluid have been made. However, the focus is now
on ideal gases. The thermal equation of state obeys

pv = RT. (12.101)

Mathematical reformulation leads to


R
p=T . (12.102)
v
The partial derivative follows accordingly:
 
∂p R
= . (12.103)
∂T v v

The combination of Eqs. 12.77, 12.101 and 12.103 results in



R
du = cv (v, T ) · dT + T − p ·dv (12.104)
v
  
=0

respectively
du = cv (v, T ) · dT (12.105)

Obviously, the specific internal energy u for ideal gases is solely a function of tem-
perature T , see Example 12.1. With reference to Eq. 12.8, the specific heat capacity
12.3 Derivation of the Caloric Equations for Real Fluids 229

cv is also only a function of temperature. Therefore, the initial Eq. 12.56 for ideal
gases simplifies to
du = cv (T ) dT (12.106)

The thermal equation of state can also be given in the following notation

RT
v= . (12.107)
p

Hence, the partial derivative follows accordingly:


 
∂v R
= . (12.108)
∂T p p

The combination of Eqs. 12.99, 12.107 and 12.108 yields



R
dh = c p ( p, T ) · dT + −T + v ·d p (12.109)
p
  
=0

respectively
dh = c p ( p, T ) · dT (12.110)

The specific enthalpy h for ideal gases is solely a function of temperature T , cf.
Example 12.1, since
h = u + RT. (12.111)

With reference to Eq. 12.10, the specific heat capacity c p is as well only a function
of temperature. Therefore, the initial Eq. 12.80 for ideal gases simplifies to

dh = c p (T ) dT (12.112)

According to the Joule-Thomson experiment for ideal gases, cf. Example 12.1, both
the specific internal energy and the specific enthalpy are pure functions of tempera-
ture. For the specific entropy it follows, see Eqs. 12.78 and 12.100:
 
δq + δψ cv ∂p cv R
ds = = · dT + · dv = · dT + · dv (12.113)
T T ∂T v T v

and
 
δq + δψ cp ∂v cp R
ds = = · dT − · dp = · dT − · d p. (12.114)
T T ∂T p T p
230 12 Caloric Equations of State

The integration leads to:

T2 v2
Δs = s2 − s1 = cv ln + R ln (12.115)
T1 v1

and
T2 p2
Δs = s2 − s1 = c p ln − R ln (12.116)
T1 p1

Consequences for incompressible liquids

Part I covers not only ideal gases, but incompressible liquids as well. According to
[5], these idealised liquids follow the simple thermal equation of state

v ( p, T ) = const. (12.117)

respectively
dv = 0. (12.118)

The simplification is that the specific volume does neither depend on the pressure
nor on the temperature, i.e.  
∂v
=0 (12.119)
∂p T

respectively  
∂v
= 0. (12.120)
∂T p

What consequences does such a simplification have for the specific internal energy
u and the specific enthalpy h?
• Specific internal energy u

Following Eq. 12.77 it is


 
∂p
du = cv (v, T ) · dT + T − p · dv = cv (v, T ) · dT. (12.121)
∂T v

Hence, the specific internal energy u follows:

du = cv (v, T ) · dT (12.122)

• Specific enthalpy h

Following Eq. 12.99 it is


12.3 Derivation of the Caloric Equations for Real Fluids 231
   
∂v
dh = c p ( p, T ) · dT + −T + v · d p = c p ( p, T ) · dT + v · d p.
∂T p
(12.123)
Since dh is the differential of a state function, cf. [5] and Eq. 10.19, it is
   
∂c p ∂v
= = 0. (12.124)
∂p T ∂T p

Thus, the specific heat capacity depends only on T , i.e.

c p = c p (T ) . (12.125)

Furthermore, the specific enthalpy h is by definition, see Sect. 11.3.1:

h = u + pv. (12.126)

The total differential for an incompressible liquid, i.e. dv = 0, follows accordingly

dh = du + p dv + v d p = du + v d p. (12.127)

In combination with Eq. 12.122 it is

dh = cv (v, T ) · dT + v d p. (12.128)

A comparison of Eqs. 12.123, 12.125 and 12.128 results in

c p ( p, T ) = c p (T ) = cv (v, T ) = cv (T ) . (12.129)

In other words, the specific heat capacity cv is also only a function of T . Further-
more, for incompressible liquids it is:

c (T ) = c p (T ) = cv (T ) (12.130)

In summary, the following can be stated for incompressible liquids:

du = c (T ) · dT (12.131)

and
dh = c (T ) · dT + v · d p (12.132)

The conclusion that the specific entropy s is a state value does not depend on
the type of fluid. It must therefore be possible to find a caloric equation of state to
calculate the state value specific entropy s also for incompressible liquids, i.e. dv = 0
as well as cv = c p = c. Thus, Eq. 12.30 simplifies to
232 12 Caloric Equations of State

δq + δψ du dT
= =c . (12.133)
T T T

The integration of the right hand side is path-independent,14 since15

2
dT T2
c = c ln . (12.134)
T T1
1

Consequently, the term must be a state and not a process value. However, this is the
newly introduced specific entropy s:

δq δψ dT
ds = + =c . (12.135)
T T T
Its integration leads to the caloric equation of state for the specific entropy of incom-
pressible liquids:
T2
Δs = s2 − s1 = c ln (12.136)
T1

Note that solids, as long as they are incompressible, can be treated with the same
means as the recently introduced incompressible liquids.

12.4 Handling of the Caloric State Equations

Initially, the aim was to find correlations for Δu = u 2 − u 1 and Δh = h 2 − h 1 in


order to solve the first law of thermodynamics for closed systems

q12 + w12 = u 2 − u 1 + Δea12 (12.137)

as well as for open systems

q12 + wt12 = h 2 − h 1 + Δea12 . (12.138)

In the last section, the caloric equations of state for entropy, enthalpy and entropy
have been derived. Now the technical application of these equations of state is pre-
sented.

14 No path information is required to solve the integral.


15 A constant specific heat capacity is assumed here. How to proceed if the specific heat capacity
is not constant is shown in Sect. 12.4.4.
12.4 Handling of the Caloric State Equations 233

12.4.1 Ideal Gases

As already mentioned, Δu and Δh cannot be measured, so they have to be calculated.


Integration of the caloric equations of state for ideal gases (12.106) and (12.112) leads
to16 :
2 2
du = u 2 − u 1 = cv dT = cv (T2 − T1 ) (12.139)
1 1

and
2 2
dh = h 2 − h 1 = c p dT = c p (T2 − T1 ) . (12.140)
1 1

Under these conditions, the first law of thermodynamics simplifies to


• Closed system
q12 + w12 = cv (T2 − T1 ) + Δea12 (12.141)

• Open system
q12 + wt12 = c p (T2 − T1 ) + Δea12 . (12.142)

12.4.2 Distinction Between cv and c p for Ideal Gases

Both cv and c p are specific heat capacities. The specific heat capacity cv indicates the
amount by which the internal energy U of a fluid rises when the temperature of 1 kg
increases by 1 K. However, c p is a measure of the amount by which the enthalpy H
increases under the same conditions. Thus, both heat capacities are thermal properties
of the fluid. In deriving the caloric equations of state, cv has been introduced as
the specific heat capacity at constant specific volume17 and c p as the specific heat
capacity with constant pressure.18 A handy explanation of the distinction between
the two specific heat capacities is shown in Example 12.2.
Theorem 12.4 No matter what change of state the system undergoes, it is always
c p that is needed to calculate Δh and always cv that is needed for Δu:

Δh → c p

respectively
Δu → cv .

16 Assuming that cv = const. and c p = const.


17 This is indicated by the index v.
18 This is indicated by the index p.
234 12 Caloric Equations of State

The definition of the specific enthalpy h is

h = u + pv. (12.143)

With the thermal equation for ideal gases it results in

h = u + pv = u + RT. (12.144)

Hence, the derivative with respect to temperature follows

dh du
= + R. (12.145)
dT dT
Following Eqs. 12.106 and 12.112 it is

du = cv (T ) dT (12.146)

and
dh = c p (T ) dT. (12.147)

Thus, Eq. 12.145 simplifies for ideal gases to

c p (T ) = cv (T ) + R (12.148)

Example 12.2 Obviously, according to Eq. 12.148 the specific heat capacity with
constant pressure c p is always greater than the specific heat capacity with constant
volume. This is verified with the following example, cf. Fig. 12.2. Both closed sys-
tems are filled with the same type and mass m of an ideal gas. System A is closed
with a horizontally movable piston so that the fluid inside is always in mechanical
equilibrium with the environment, i.e. the pressure of the fluid inside the cylinder is
always equal to the ambient pressure pA = penv . Hence, the change of state is iso-
baric. According to Example 7.6, only for very slow changes of state19 (A = 0) and
a non-accelerated (a = 0),20 frictionless (fric. = 0) piston the process is isobaric.
Thus, case A needs to be reversible.

System B is a rigid vessel so that its volume is constant, i.e. the change of state
is isochoric. Initially, the temperature in both systems is the same. Now, the same
amount of heat, i.e. Q 12 , is supplied to both systems. The question is, how the
temperature rise differs in both systems:

19 Very slow is a synonym for no turbulence inside, see also Theorem 7.16.
20 The acceleration due to gravity g is not relevant because the piston is operated horizontally.
12.4 Handling of the Caloric State Equations 235

Fig. 12.2 Isobaric (Case A) versus isochoric (Case B) change of state

• Case A: Isobaric change of state ( p = const.)


First law of thermodynamics in differential notation21 :

δ Q + δWA = dUA . (12.149)



<0

Obviously, work δWA is released: Due to the heat supply, the gas inside the cylin-
der heats up and expands while the pressure remains constant.22 This expansion
is the cause of the system releasing work.

Partial energy equation:

δWA = δWV,A + δA = − pA dV. (12.150)



=0

Hence, the first law of thermodynamics obeys

δ Q = dUA + pA dV. (12.151)

The enthalpy H is defined as:

H = U + pV. (12.152)

Its differential for an isobaric, i.e. d p = 0, change of state is:

dH = dU + p dV + V d p = dU + p dV. (12.153)

Thus, for case A it obviously is

dUA = dHA − pA dV. (12.154)

21 The change in kinetic energy can be neglected because the change of state is quasi-static. Fur-
thermore, there is no change in potential energy because the cylinder is horizontal.
22 According to V = m RT .
p
236 12 Caloric Equations of State

According to Eq. 12.151 this means:

δ Q = dHA . (12.155)

In combination with the caloric equation of state for the enthalpy23 it results in

δ Q = mc p dTA . (12.156)

Integration leads to
Q 12 = mc p ΔTA . (12.157)

In other words:
Q 12
cp = . (12.158)
m ΔTA

• Case B: Isochoric change of state (v = const.)


The first law of thermodynamics in differential notation yields

δ Q + δWB = dUB . (12.159)

Partial energy equation

δWB = 0 = − pB dV + δB . (12.160)

Obviously, according to Fig. 12.2, no work δWB passes the system boundary. Fur-
thermore, the volume does not vary, i.e. dV = 0. Consequently, there can not be
any dissipation, i.e. δB = 0. That makes sense because there is no fluid movement
inside the cylinder.24 Thus, the first law of thermodynamics simplifies to

δ Q = dUB . (12.161)

In combination with the caloric equation for the internal energy25 it results in

δ Q = mcc dTB . (12.162)

Integration leads to
Q 12 = mcv ΔTB . (12.163)

In other words:
Q 12
cv = . (12.164)
m ΔTB

23 Note that H = mh.


24 Fluid movement would have to be triggered by a moving piston, for example.
25 Note that U = mu.
12.4 Handling of the Caloric State Equations 237

Let us now compare Eqs. 12.158 and 12.164: It is obvious that the temperature
increase is greater in case B than in case A. This is a consequence of the first law of
thermodynamics, cf. Eq. 12.149 respectively Eq. 12.161: In case B, all the supplied
heat leads to an increase in internal energy, whereas in case A, part of the sup-
plied energy is released as work. Consequently, the remaining energy to increase the
internal energy is less than in case B. The ratio of Eqs. 12.158 and 12.164 results in

cp ΔTB
= >1 (12.165)
cv ΔTA

Due to c p > cv the temperature rise of B must be greater than the temperature increase
of A. The specific heat capacities thus indicate how much energy is needed to raise
the temperature of 1 kg of a fluid by 1 K. This depends, as this example has shown,
on the process control.
Problem 12.1 This problem refers to case A according to Fig. 12.2. The cylinder
is filled with m = 1 kg of air (ideal gas, R = 287 kgJK , cv = 717 kgJK ). The initial
temperature is T1 = 300 K, the pressure is supposed to be p = 1 bar = const. while
the system is heated up by Q = 10 kJ. The entire change of state shall be reversible.
(a) Calculate the initial volume V1 .
(b) Determine the final temperature T2 .
(c) What is volume V2 ?
(d) How much work W12 is released?
(e) Sketch the change of state (1) → (2) in a p, V - as well as in a T, S-diagram.26
Solution
(a) The thermal equation of state is applied for the calculation of the initial volume
V1 :
m RT1
V1 = = 0.861 m3 . (12.166)
p

(b) The first law of thermodynamics obeys

Q 12 + W12 = U2 − U1 . (12.167)

The partial energy equation for W12 results in

2
W12 = WV,12 + 12 = − p dV. (12.168)

=0 1

There is no dissipation because the change of state shall be reversible. Since the
change of state is also isobaric, the volume work can be easily calculated:

26This part is intended for advanced readers who are already familiar with the T, s-diagram, see
Sect. 13.4.
238 12 Caloric Equations of State

W12 = − p ΔV = − p (V2 − V1 ) . (12.169)

Hence, the first law of thermodynamics follows

Q 12 = U2 − U1 + p (V2 − V1 ) = H2 − H1 = m (h 2 − h 1 ) . (12.170)

Applying the caloric equation of state (12.140) leads to27

Q 12 = mc p (T2 − T1 ) (12.171)

with
J
c p = cv + R = 1004 . (12.172)
kg K

Temperature T2 follows

Q 12
T2 = T1 + = 309.96 K. (12.173)
mc p

(c) The thermal equation of state results in

m RT2
V2 = = 0.8896 m3 . (12.174)
p

(d) According to Eq. 12.169 it is

W12 = − p ΔV = − p (V2 − V1 ) = −2.8586 kJ. (12.175)

Obviously, of the 10 kJ thermal energy supplied, 2.8586 kJ are released as


mechanical work. Thus 7.1414 kJ are utilised to increase the internal energy
of the air in the cylinder.
(e) The change of state (1) → (2) is visualised in a p, V - and a T, S-diagram, see
Fig. 12.3.

12.4.3 Isentropic Exponent

The so-called isentropic exponent is defined as

cp
κ= . (12.176)
cv

27 Assuming, that c p = const.


12.4 Handling of the Caloric State Equations 239

Fig. 12.3 Solution to Problem 12.1

Table 12.1 Isentropic exponent for ideal gases at standard conditions, i.e. temperature ϑ = 0 ◦ C
according to DIN 1343, see [1]
Number of atoms κ
1 1.67
2 1.40
3 1.33

This exponent has an important meaning when the isentropic change of state is
introduced, cf. Sect. 13.5. In any case, it makes sense to define κ already at this
stage. For ideal gases, the isentropic exponent as a function of the number of atoms
of which they are composed is given in Table 12.1.
Applying Eq. 12.148 and the definition of the isentropic exponent leads to

c p − cv = R (12.177)

and κ
cp = R (12.178)
κ −1

and
1
cv = R. (12.179)
κ −1

12.4.4 Temperature Dependent Specific Heat Capacity

For ideal gases, the specific heat capacities cv and c p are pure functions of the
temperature ϑ, see Sect. 12.4.2. The following correlation between c p and cv has
been derived:
c p (ϑ) = R + cv (ϑ) . (12.180)
240 12 Caloric Equations of State

For small temperature differences, the temperature dependence of the specific heat
capacities can often be neglected. However, in this section the temperature depen-
dency is considered—this becomes relevant for significant temperature changes.
Graphite, for example, shows a strong temperature dependence, while copper remains
almost constant in the temperature range between 100 . . . 1000 ◦ C, see [23].

Arithmetic averaged specific heat capacity

The caloric equations of state for specific internal energy as well as for specific
enthalpy obey
2
du = cv dT ⇒ u 2 − u 1 = cv (ϑ) dϑ (12.181)
1

respectively
2
dh = c p dT ⇒ h 2 − h 1 = c p (ϑ) dϑ. (12.182)
1

It does not make any difference which temperature scale (◦ C or K) is applied, since

ϑ = T − 273.15 K ⇒ dϑ = dT. (12.183)

Instead of denoting the temperature dependence for cv and c p separately, c represents


both specific heat capacities in this section. Figure 12.4 illustrates qualitatively how
the specific heat capacity c correlates with temperature ϑ. With rising temperature,
the specific heat capacity increases as well. The goal is now to find a solution for the
integrals in Eqs. 12.181 and 12.182. Thus, an averaged specific heat capacity c|ϑϑ21 is
introduced in the temperature range of ϑ1 and ϑ2 :

Fig. 12.4 Temperature


dependency of the specific
heat capacity
12.4 Handling of the Caloric State Equations 241

2
c(ϑ) dϑ = c|ϑϑ21 (ϑ2 − ϑ1 ). (12.184)
1

Thus, the arithmetic averaged specific heat capacity yields

2
1
c|ϑϑ21 = · c(ϑ) dϑ. (12.185)
(ϑ2 − ϑ1 )
1

In tables, e.g. in [24], averaged values are usually given. The reference value in these
tables is usually 0 ◦ C, which simplifies the calculation:


1
c|ϑ0 = · c(ϑ) dϑ. (12.186)
ϑ
0

It follows
ϑ1
ϑ1 · c|ϑ0 1 = c(ϑ) dϑ (12.187)
0

respectively
ϑ2
ϑ2 · c|ϑ0 2 = c(ϑ) dϑ. (12.188)
0

For the integrals it is, see Fig. 12.4

ϑ2 ϑ2 ϑ1


c(ϑ) dϑ = c(ϑ) dϑ − c(ϑ) dϑ. (12.189)
ϑ1 0 0

Thus, it finally results in

ϑ2  
c(ϑ) dϑ = c|ϑ0 2 · ϑ2 − c|ϑ0 1 · ϑ1 (12.190)
ϑ1

respectively
c|ϑ0 2 · ϑ2 − c|ϑ0 1 · ϑ1
c|ϑϑ21 = (12.191)
ϑ2 − ϑ1
242 12 Caloric Equations of State

Logarithmic averaged specific heat capacity

For the caloric state value specific entropy, the averaging of the specific heat capac-
ity differs from the procedure for specific internal energy or specific enthalpy.
Section 12.3 has shown, that

dT dp
ds = c p −R (12.192)
T p

so that

T p
dT dp ϑ T p
s − s0 = cp − R = c p ϑ0 ln − R ln . (12.193)
T p T0 p0
T0 p0

This leads to the logarithmic averaged specific heat capacity

T
ϑ 1 dT
c p ϑ0 = T cp (12.194)
ln T0 T
T0

Note that s0 = s (ϑ0 , T0 , p0 ) is a reference level. The entropy difference between


two states (1) and (2) then reads as
ϑ T2 p2
s2 − s0 = c p ϑ20 ln − R ln (12.195)
T0 p0

and
ϑ T1 p1
s1 − s0 = c p ϑ10 ln − R ln . (12.196)
T0 p0

Thus, it is
   
ϑ T2 p2 ϑ T1 p1
s2 − s1 = c p ϑ20 ln − R ln − c p ϑ10 ln − R ln
T0 p0 T0 p0
(12.197)
ϑ2 T2 ϑ1 T1 p  ϑ T p2
= c p ϑ0 ln − c p ϑ0 ln − R ln ≡ c p ϑ21 ln − R ln .
2 2
T0 T0 p1 T1 p1

Solving for the logarithmic averaged specific heat capacity in the temperature range
T1 . . . T2 leads to
ϑ ϑ
ϑ2 c p ϑ20 ln TT20 − c p ϑ10 ln TT01
c p ϑ1 = (12.198)
ln TT21
12.4 Handling of the Caloric State Equations 243

Alternatively, the specific entropy can be calculated according to Sect. 12.3 as


follows
dT dv
ds = cv +R (12.199)
T v
so that

T v
dT dv ϑ T v
s − s0 = cv + R = cv ϑ0 ln + R ln . (12.200)
T v T0 v0
T0 v0

This leads to the logarithmic averaged specific heat capacity

T
ϑ 1 dT
cv ϑ0 = T cv (12.201)
ln T0 T
T0

Note that s0 = s (ϑ0 , T0 , v0 ) is a reference level. The entropy difference between


two states (1) and (2) then reads as
ϑ T2 v2
s2 − s0 = cv ϑ20 ln + R ln (12.202)
T0 v0

and
ϑ T1 v1
s1 − s0 = cv ϑ10 ln + R ln . (12.203)
T0 v0

Thus, it is
   
ϑ T2 v2 ϑ T1 v1
s2 − s1 = cv ϑ20 ln + R ln − cv ϑ10 ln + R ln
T0 v0 T0 v0
(12.204)
ϑ2 T2 ϑ1 T1 v  ϑ T v2
= cv ϑ0 ln − cv ϑ0 ln + R ln ≡ cv ϑ21 ln + R ln .
2 2
T0 T0 v1 T1 v1

Solving for the logarithmic averaged specific heat capacity in the temperature range
T1 . . . T2 leads to
ϑ ϑ
ϑ2 cv ϑ20 ln TT02 − cv ϑ10 ln TT01
cv ϑ1 = (12.205)
ln TT21

A more detailed explanation on averaged specific heat capacities28 is shown in


Sect. B.

28 Including arithmetic and logarithmic averaged values.


244 12 Caloric Equations of State

Table 12.2 Temperature dependence of an ideal gas according to Problem 12.2


ϑ [◦ C] 200 1200
 
ϑ
c p |0 kg K
kJ
1.052 1.242

Problem 12.2 The specific heat capacity at constant pressure c p of an ideal gas is
temperature dependent and follows Table 12.2. Please calculate the arithmetic aver-
aged specific heat capacity at constant pressure within the temperatures ϑ1 = 200 ◦ C
and ϑ2 = 1200 ◦ C. What is Δh = h 2 − h 1 ?

Solution

The averaged heat capacity follows Eq. 12.191, so that

c p |ϑ0 2 · ϑ2 − c p |ϑ0 1 · ϑ1
c p |ϑϑ21 = (12.206)
ϑ2 − ϑ1

1.242 · 1200 − 1.052 · 200 kJ kJ


⇒ c p |ϑϑ21 = = 1.28 . (12.207)
1200 − 200 kg K kg K

For the difference of specific enthalpy it follows

2
h2 − h1 = c p (ϑ) dϑ = c p |ϑϑ21 (ϑ2 − ϑ1 ) (12.208)
1

respectively
kJ
h 2 − h 1 = c p |ϑϑ21 (ϑ2 − ϑ1 ) = 1280 . (12.209)
kg

12.4.5 Incompressible Fluids, Solids

Note that c = c p = cv for incompressible liquids, i.e. it is not required to distinguish


between them. Solids are also treated as incompressible in Part I, so that c = c p = cv
also applies here. Thus, the caloric equations of state follow, see Sect. 12.3.2:
• Specific internal energy u
du = c (ϑ) dϑ (12.210)

Integration leads to29

29 A temperature difference is needed for the calculation, so it makes no difference whether one
takes Δϑ or ΔT , see Eq. 12.183.
12.4 Handling of the Caloric State Equations 245

u 2 − u 1 = c|ϑϑ21 (ϑ2 − ϑ1 ) (12.211)

• Specific enthalpy h
dh = c (ϑ) dϑ + v d p (12.212)

Since v = const., integration leads to

h 2 − h 1 = c|ϑϑ21 (ϑ2 − ϑ1 ) + v ( p2 − p1 ) (12.213)

Thus, the term v ( p2 − p1 ) is additionally relevant in case the change of state from
(1) to (2) is non-isobaric.
• Specific entropy s
dT
ds = c (ϑ) (12.214)
T
Thus, it is
ϑ T2
s2 − s1 = cϑ21 ln (12.215)
T1

In case the temperature dependence is irrelevant this equation simplifies to

T2
s2 − s1 = c ln . (12.216)
T1

Note that c represents the arithmetic averaged specific heat capacity, whereas c is
the logarithmic averaged specific heat capacity, cf. Sect. 12.4.4.
Problem 12.3 Two identical vertical pipes with a diameter of D = 0.1 m are con-
nected by a thin tube and a valve, see Fig. 2.8. One pipe is filled with water up to a
height of H = 10 m. Water has a density of ρ = 1000 mkg3 , the other tube is initially
empty. Now the valve is opened. After some time, an equilibrium of the water quan-
tity is reached. What is the temperature rise of the water? Water shall be treated as
an incompressible liquid with c = 4.18 kgkJK .

Solution

The rise of internal energy ΔU has previously been calculated in Problem 11.2 by
applying the first law of thermodynamics for closed systems:

H
ΔU = mg = 1926.2J. (12.217)
4
In order to calculate the temperature increase, the caloric equation of state has to be
applied:
ΔU = m Δu = mc ΔT. (12.218)
246 12 Caloric Equations of State

Hence, it is

ΔU mg H4 gH
ΔT = = = = 5.8672 × 10−3 K. (12.219)
mc mc 4c
The problem has even been calculated dynamically, see Problem 11.5. Figure 11.19
shows the same temperature rise ΔT once the system comes to rest.

Problem 12.4 Steam must be condensed in a heat exchanger of a thermal power


plant. Therefore, a thermal power of Q̇ = 75 × 103 kW has to be transferred. For
this purpose, cooling water is taken from a river that flows by with a total mass flux
of ṁ total = 20 × 103 kgs , see Fig. 12.5. The system is regarded in steady state.
(a) What mass flow rate ṁ cond. has to be taken from the river, if a temperature rise
of ΔTcond. = 3 K of the coolant stream is tolerated?
(b) What temperature rise ΔTtotal results for the river after complete mixing of the
coolant flow, if this can be regarded as adiabatic?
Hint: There is to be no pressure loss in the condenser. Change of outer energies can
be ignored, the specific heat capacity of water is cW = 4.19 kgkJK .

Solution
(a) To calculate the mass flow rate, the first law of thermodynamics for open systems
is applied according to the system boundary A, see Fig. 12.5.
With reference to Sect. 11.3.3 the energy flux into the system needs to be bal-
anced by the energy flux out of the system in steady state. Thus, the first law of
thermodynamics results in

Fig. 12.5 Sketch to Problem 12.4


12.4 Handling of the Caloric State Equations 247

ṁ cond. h 1 + Q̇ = ṁ cond. h 2 . (12.220)

In order to apply the caloric equation of state a difference of enthalpies is required,


so the first law of thermodynamics needs to be reformulated:

Q̇ = ṁ cond. (h 2 − h 1 ) . (12.221)

Applying the caloric equation of state for an incompressible liquid without pres-
sure loss
h 2 − h 1 = cW (T2 − T1 ) + v ( p2 − p1 ) (12.222)
  
=0

yields
Q̇ = ṁ cond. (h 2 − h 1 ) = ṁ cond. cW (T2 − T1 ) . (12.223)
  
=3 K

Thus, the mass flow rate is

Q̇ kg
ṁ cond. = = 5.967 × 103 . (12.224)
cW (T2 − T1 ) s

(b) To calculate the temperature increase of the flux ΔTtotal = T3 − T1 , the first law
of thermodynamics is applied for system boundary B:

ṁ total h 1 + Q̇ = ṁ total h 3 . (12.225)

Rearranging leads to
Q̇ = ṁ total (h 3 − h 1 ) . (12.226)

Applying the caloric equation of state for isobaric, incompressible liquids results
in
Q̇ = ṁ total cW (T3 − T1 ) . (12.227)

Thus, the temperature rise of the river follows


ΔTtotal = T3 − T1 = = 0.895 K. (12.228)
ṁ total cW

Alternatively, the first law of thermodynamics can be applied for system bound-
ary C. Steady state balancing of incoming and outgoing energy fluxes results in

(ṁ total − ṁ cond. ) h 1 + ṁ cond. h 2 = ṁ total h 3 . (12.229)

For applying the caloric equations of state, the first law of thermodynamics has
to be rearranged for enthalpy differences, i.e.
248 12 Caloric Equations of State

ṁ total (h 3 − h 1 ) = ṁ cond. (h 2 − h 1 ) . (12.230)

Substituting the caloric equations of state yields

ṁ total cW (T3 − T1 ) = ṁ cond. cW . (T2 − T1 ) (12.231)


     
=ΔTtotal =ΔTcond.

Thus, it finally is
ΔTcond. ṁ cond.
ΔTtotal = = 0.895 K. (12.232)
ṁ total

Problem 12.5 Air (T1 = 300 K and p1 = 1000 kPa) passes an adiabatic throttle.
Thereby, its pressure changes to p2 = 700 kPa. The inlet velocity is c1 = 20 ms . The
cross sections of the tube at inlet A1 and outlet A2 shall be the same. Air can be
treated as an ideal gas (cv = 0.717 kgkJK , R = 287 kgJK ). What is the temperature T2
after throttling? The change of kinetic energy shall not be ignored!

Solution

Let us start with the first law of thermodynamics for this steady state problem, see
Fig. 12.6. The energy flux into the system needs to be balanced by the energy flux
out of the system, otherwise the problem could not be steady state:

1 2 1 2
ṁ 1 h 1 + ṁ 1 c + gz 1 + Q̇ 12 + Pt12 = ṁ 2 h 2 + ṁ 2 c + gz 2 . (12.233)
2 1   2 2
=0 =0

The mass balance at steady state means that the mass flow into the system is equal
to the mass flow out of the system, so that

ṁ 1 = ṁ 2 . (12.234)

Since no information regarding the change of potential energy is given, it is ignored.


Hence, the first law of thermodynamics simplifies to

Fig. 12.6 Sketch to Problem 12.5


12.4 Handling of the Caloric State Equations 249

1 2

c2 − c12 + (h 2 − h 1 ) = 0. (12.235)
2
Applying the caloric equation of state:

1 2

c2 − c12 + c p (T2 − T1 ) = 0 (12.236)


2
with
J
c p = cv + R = 1004 . (12.237)
kg K

In order to gather information regarding the velocities, the equation of continuity is


applied, i.e.
ṁ 1 = ṁ 2 ⇒ ρ1 c1 A1 = ρ2 c2 A2 . (12.238)

Hence, it is
ρ1 v2
c2 = c1 = c1 . (12.239)
ρ2 v1

The fluid is an ideal gas, so that the thermal equation of state can be applied

RT1
p1 v1 = RT1 ⇒ v1 = (12.240)
p1

and
RT2
p2 v2 = RT2 ⇒ v2 = . (12.241)
p2

Consequently, the outlet velocity is

p1 T2
c2 = c1 . (12.242)
p2 T1

Substituting in the first law of thermodynamics (12.235) results in


  2 
1 p 1 T2
T2 − T1 = − c2 − c12 (12.243)
2c p 1 p2 T1

 2
c12 p1 c12
⇒ T22 + T2 − + T1 = 0. (12.244)
2c p p2 T1 2c p

This quadratic equation has two solutions:

T2,1 = 299.7932 K (12.245)


250 12 Caloric Equations of State

respectively
T2,2 = −2.2168 × 105 K. (12.246)

Thus, only solution T2,1 = 299.7932 K makes sense. The corresponding velocity
follows Eq. 12.242 and leads to c2 = 28.55 ms .

12.4.6 Adiabatic Throttle

As mentioned earlier, see Problem 12.5, a throttle is a simple work-insulated com-


ponent used to reduce the pressure of a fluid. The principle sketch of an adiabatic
throttle is shown in Fig. 12.6. In the context of Problem 12.5, this component has
already been calculated in detail, but now relevant assumptions are made:
In steady state the equation of continuity reads as:

ṁ 1 = ṁ 2 = ṁ. (12.247)

The first law of thermodynamics for an adiabatic throttle, i.e. an work-insulated


component, in steady state obeys

1 2 1 2
ṁh 1 + ṁ c1 + gz 1 + Q̇ 12 + Pt12 = ṁh 2 + ṁ c2 + gz 2 . (12.248)
2   2
=0 =0

After reformulation to a difference notation, the following results



1 2

0 = ṁ h 2 − h 1 + c − c1 + g (z 2 − z 1 ) .
2
(12.249)
2 2

Since the dimensions of such a component are rather small, the change in potential
energy can be neglected. Hence, the first law of thermodynamics further simplifies to

1 2

h2 − h1 + c2 − c12 = 0. (12.250)
2

If one also disregards the kinetic energy, the following results from the first law of
thermodynamics
h2 − h1 = 0 (12.251)

Consequently, the change of state is called isenthalpic.30 If the fluid is an ideal gas,
the specific enthalpy is a pure function of temperature, i.e.

30 The specific enthalpy remains constant.


12.4 Handling of the Caloric State Equations 251

c p (T2 − T1 ) = 0 ⇒ T2 = T1 . (12.252)

The assumption of neglecting the kinetic energy is sufficiently accurate in many


cases, as Problem 12.5 has shown. The conclusion from this problem was that the
inlet and outlet temperatures are approximately the same, i.e. the kinetic energy does
not make a significant contribution.
Theorem 12.5 Neglecting the change in outer energies, the change of state in an
adiabatic throttle is isenthalpic. If an adiabatic throttle is operated with an ideal gas,
the change of state is isothermal.
However, if it is a real fluid, see Part II, i.e. a non-ideal fluid, the specific enthalpy
is not exclusively a function of temperature but also of pressure. Thus, a tempera-
ture change may be measured while the real fluid is throttled. This so-called Joule-
Thomson effect is dealt with in Part II, see Sect. 18.5.

How is the pressure loss in a throttle technically achieved?

Let us have a look at the partial energy equation. Since the component is a work-
insulated system, it obeys

2
1 2

wt = 0 = v d p + ψ12 + c − c12 + g (z 2 − z 1 ) . (12.253)


2 2
1

As discussed before, the outer energies are ignored, so that it simplifies to:

2
v d p = −ψ12 ≤ 0 . (12.254)
1

Hence, it finally is
d p ≤ 0. (12.255)

Theorem 12.6 Obviously, due to dissipation within the throttle the fluid’s pressure
decreases.
Chapter 13
Meaning and Handling of Entropy

In the previous Chap. 12, the caloric equations of state for internal energy as well
as for enthalpy have been introduced and derived mathematically. Furthermore, a
new state value, the specific entropy s, has been introduced. At that time, however,
it has not yet been clear how entropy can be used to evaluate thermodynamic sys-
tems. In this chapter, a clarification is given as to why entropy is useful and why
it is indispensable for describing energetic transformation. First, a comparison with
the first law of thermodynamics is made so that entropy balancing becomes compre-
hensible: Unlike energy, entropy is not a conservation variable, since entropy can be
generated in a system. Nevertheless, entropy can be balanced. A distinction is made
between closed/open systems and thermodynamic cycles. In addition, a new state
diagram, the T, s-diagram, is derived. Such a state diagram visualises the process
values specific heat q and specific dissipation ψ. Together with the p, v-diagram,
which visualises the other process value, namely the specific work, it is obviously
an important diagram in thermodynamics. Finally, two new changes of state, namely
isentropic and polytropic change of state, are treated in this chapter.

13.1 Entropy—Clarification

Let us take a closer look at what is known about entropy so far. According to Eq. 12.44
it is
δq δψ
ds = + . (13.1)
T T
 
=δsa =δsi

Hence, the state value specific entropy s of a fluid can be modified by two mecha-
nisms, see Fig. 13.1:

© Springer Nature Switzerland AG 2022 253


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_13
254 13 Meaning and Handling of Entropy

Fig. 13.1 Mechanism of entropy in a closed systems

1. Specific heat that passes the system boundary always carries specific entropy with
it. Obviously, this part of the specific entropy is supplied from the environment,
or released to the environment. This part follows

δq
δsa = , (13.2)
T
integration leads to:
2 2
δq
δsa = = sa,12 . (13.3)
T
1 1

If no specific heat exceeds the system boundary, i.e. the system is adiabatic, there
is also no transport of specific entropy across the system boundary. In addition to
the quantity of specific heat q, the temperature T inside the system at the location
where the heat transfer takes place is obviously also relevant.
2. The second mechanism for influencing the specific entropy of a system takes
place within the system, i.e.
δψ
δsi = , (13.4)
T
integration leads to:
2 2
δψ
δsi = = si,12 . (13.5)
T
1 1
13.1 Entropy—Clarification 255

Unlike δsa , which can have any sign, δsi can only be positive, or at best become
zero, because of ψ ≥ 0. The best case would be a reversible change of state, i.e.
ψ = 0. Note that T in Eq. 13.4 is the temperature in the system at the location
where dissipation occurs. In case the system is homogeneous, see Sect. 3.1.1,
the temperature T for δsa is the same as for δsi . Homogeneous systems have a
uniform temperature.1
The integration of Eq. 13.1 reads as

2
ds = s2 − s1 = sa,12 + si,12 (13.6)
1

However, according to Eq. 13.6 the change of entropy can be positive, negative or it
can become zero:
• In case specific heat is supplied and specific dissipation occurs, the specific entropy
inside the system rises.
• In case the system is adiabatic, the specific entropy can be constant or it can rise.
However, it can not sink.
• In case the system is cooled, the specific entropy can increase, decrease or remain
constant. It depends on the ratio of the specific heat released to the specific dissi-
pation.
The extensive entropy2 results in
S = ms (13.7)

with
J
[S] = 1 (13.8)
K
respectively
J
[s] = 1 . (13.9)
K kg

13.2 Comparison Entropy Balance Versus First Law


of Thermodynamics

In Fig. 13.2 the balance of entropy for a closed system is illustrated. Note that exten-
sive values should always be applied for thermodynamic balances. In state (1) the

1 Depending on the system boundary, it is usually assumed that the temperature at which heat
transfer occurs is the same as that of dissipation.
2 Analogous to the first law of thermodynamics, entropy balances should always be conducted

extensively.
256 13 Meaning and Handling of Entropy

Fig. 13.2 Balance of entropy in a closed system

system can be described by its state values, such as pressure p1 and temperature
T1 for instance. In addition to these intensive state values, there are also extensive
state values, e.g. U1 and H1 . Anyhow, as derived in Chap. 12, the extensive entropy
S1 = ms1 is a state value as well. The system thus contains entropy S1 in state (1),
which is shown as bubbles in Fig. 13.2. Due to external influences, however, the
system can be transferred to a new state of equilibrium (2), i.e. depending on the
change of state, some state values have changed. Hence, the amount of entropy may
have changed as well. The mechanisms how entropy varies have been discussed
already: Due to heat transfer Q 12 entropy Sa,12 is carried across the system boundary.
Additionally, there may be dissipation inside the system. Dissipation, however, is a
source term for entropy,3 i.e. Si,12 is generated inside the system. Both mechanisms,
heat transfer from or to the environment, as well as the generated entropy, are quan-
tified by entropy bubbles, cf. Fig. 13.2. Consequently, an entropy balance according
to Fig. 13.2 corresponds to counting bubbles, i.e.

S2 = S1 + Sa,12 + Si,12 . (13.10)

The entropy in state (2) is the initial amount of entropy in state (1) plus entropy
crossing the system boundary4 and plus entropy, that is generated internally. Both,
Sa,12 and Si,12 , depend on the process, i.e. are process values. In other words the
entropy balance obeys
S2 − S1 = Sa,12 + Si,12 (13.11)

3 As already mentioned, entropy is no conservation value as for example energy.


4 Note that in the case of entropy release by the system due to cooling, Sa,12 would be negative.
13.2 Comparison Entropy Balance Versus First Law of Thermodynamics 257

Fig. 13.3 Balance of energy in a closed system

However, the same result occurs when multiplying Eq. 13.6 with the mass of the
system m. On the left hand side of Eq. 13.11 there is a difference of state values
which is equalised by process values on the right hand side. From now on, the entropy
balance is referred to somewhat simplified as the second law of thermodynamics.5
Obviously, the second law of thermodynamics according to Eq. 13.11 has the same
structure as the first law of thermodynamics, that is illustrated in Fig. 13.3: In state
(1) the system contains a state value extensive energy6 E 1 . The external influence of
the process values heat and work on the system varies the amount of energy in the
system. Consequently, in state (2) the energy of the system is E 2 . The first law of
thermodynamics, as derived in the previous chapters, obeys

E 2 − E 1 = W12 + Q 12 (13.12)

Comparing Eqs. 13.11 and 13.12 shows the same structure: A difference of an exten-
sive state value is balanced by process values. The energy in Eq. 13.12 can be sub-
stituted, i.e.
U2 − U1 + E a,12 = W12 + Q 12 . (13.13)

By doing so, the difference of the internal energy in Eq. 13.13 can be calculated with
a caloric equation of state, e.g. for an ideal gas:

U2 − U1 = m (u 2 − u 1 ) = mcv (T2 − T1 ) . (13.14)

5 However, the actual second law of thermodynamics and the corresponding principle are introduced
in Chap. 15.
6 Energy consists of internal energy as well as mechanical energy.
258 13 Meaning and Handling of Entropy

The same can be done with Eq. 13.11. If the fluid in Fig. 13.2 is an ideal gas, the
caloric equation of state reads as follows, see Sect. 12.3.2:
   
T2 v2 T2 p2
S2 − S1 = m (s2 − s1 ) = m cv ln + R ln = m c p ln − R ln .
T1 v1 T1 p1
(13.15)
Therefore, the conclusion of this section is that both the first and second law of ther-
modynamics have the same structure: The energy respectively entropy differences
can be calculated by a balance taking into account the relevant process values in
each case. Furthermore, the caloric equations of state can be applied to calculate the
difference in internal energy as well as the difference in entropy between state (2)
and state (1).
Theorem 13.1 Entropy can be balanced, see Eq. 13.11. Unlike energy, however,
entropy is not a conservation variable, since entropy can be generated within a
system, i.e. a source term can occur.

13.3 Energy Conversion—Why Do We Need Entropy?

So far, entropy has been treated rather mathematically. The benefits of entropy have
probably not yet become clear. This section now shows why entropy is needed as a
new state value. For this reason, a so-called thermal engine, i.e. a machine that con-
verts heat into mechanical energy, is being studied. An example of such a machine is a
thermal power plant, which in a first step converts the chemically bonded energy into
thermal energy by burning the fuel. This thermal energy is converted into mechan-
ical energy in a next step, for example by a turbine. Finally, the turbine can drive a
generator that supplies electrical energy. Machines like this play an important role in
technical thermodynamics. Since not much is known about power plants so far, it is
advantageous to use a so-called black-box notation as shown in Fig. 13.4. Anyhow,
this cycle is designed as an open system inside, i.e. the components mentioned are
flowed through by a mass flow. Such cycles are discussed later in Chaps. 17 and 22.
The power plant with its components is located within the system boundary. What
can be observed are the fluxes across the system boundary without having informa-
tion about the thermodynamic cycle7 inside. In this example, the power plant is to be
operated in stationary mode, i.e. the state of the system is constant in time. In other
words, if sensors for all state values were placed everywhere in the system, they all
would send a signal that is constant in time. However, a steady state does not mean
that the signal is homogeneous within the system: For example, the pressure at each
point would be different, but it would not change over time.
Theorem 13.2 A machine in continuous and cycle operation must be in a steady
state, i.e. all state values within the system must be constant in time. All thermody-

7 However, there must be a cycle inside, as the power plant is to be in operation around the clock.
13.3 Energy Conversion—Why Do We Need Entropy? 259

Fig. 13.4 Power plant as a block-box

namic cycle processes are characterised by several changes of state that run period-
ically so that the initial state is always reached again.
Now, let us have a closer look at the power plant: A heat flux Q̇ in , released by a
combustion process,8 enters the thermal engine on top of the black-box. At that
location the process has the highest temperature due to the combustion. Anyhow,
power P is released by the thermal engine: From the management’s point of view
this amount of work should be maximised in order to make a profit. However, thermal
power plants are usually built near a river because they obviously need to be cooled.
Cooling takes place at the bottom side of the black box9 where temperature is rather
small. Thus, the thermal engine operates between a so-called hot reservoir, i.e. the
combustion chamber, and a cold reservoir which can be a river for instance. The
question now is how much power can be released in best case, i.e. the power plant
shall be free of any dissipation und thus be reversible. In order to answer this question,
the first law of thermodynamics for such a thermodynamic cycle is applied. The
law of conservation of energy for this kind of process has already been derived in
Sect. 11.2.1, see Eq. 11.14:
Theorem 13.3 The energy that is supplied to the system must be balanced by the
energy that is released:
Ė in = Ė out . (13.16)

If this is not the case, the energy inside may either increase over time, leading to
an increase in temperature, or the energy and thus the temperature inside decrease.10
Hence, according to Fig. 13.4a the first law of thermodynamics obeys

8 Represented by a hot reservoir.


9 Represented by a cold reservoir. As already know to release heat a colder environment is required.
10 Similar to a bathtub whose water level is not constant if inflow and outflow are different.
260 13 Meaning and Handling of Entropy

Q̇ in = P + Q̇ out . (13.17)

The internal energy in the system thus remains constant in time, i.e.

dU
= Q̇ in − P − Q̇ out = 0. (13.18)
dt

When studying thermodynamic cycles,11 e.g. a power plant, the direction of the
arrows pointing into and out of the system should be assumed to be as they are in
reality,12 i.e. if a heat flux is supplied to the system, the arrow should point into the
system, while if the system is cooled, it should point outwards. Absolute, i.e. positive
values, are then used in the balances. The sign is automatically taken into account by
the direction of the arrow. Solving the energy balance for the power, that is released
by the power plant, leads to
P = Q̇ in − Q̇ out . (13.19)

The initial question was, how to maximise the power P that is released. Following
Eq. 13.19 it is not wise to cool the machine, since cooling reduces the released
power. From this point of view, cooling should be avoided completely: In such a case
the entirely supplied heat flux could be converted into power. Evidently, this does
not work, as cooling takes place in every thermal power plant. This contradiction
obviously cannot be resolved by applying the first law of thermodynamics alone.
Consequently, a second equation is required, since with Q̇ out and P there are two
unknowns, while the thermal heat flux Q̇ in is to be given. Anyhow, the premise is,
that the power plant runs in steady state, so that all state values inside need to be
constant, including the entropy as newly introduced state value:
Theorem 13.4 In steady state, the entropy that is supplied to the system must be
balanced by the entropy that leaves the system:

Ṡin = Ṡout . (13.20)

This ensures that the entropy in the system remains constant, i.e.

dS
= Ṡin − Ṡout = 0. (13.21)
dt
If this were not the case, the entropy in the system could not be constant in time.
It would either rise or decrease. As a consequence, the state of the system could not
be constant by time.13 This would contradict a steady-state operation. In our case, a

11 Thermodynamic cycles have already been discussed in Sects. 8.3 and 8.4.3.
12 This is just a recommendation.
13 For example, if the entropy as a state value in the system increases, e.g. due to pure heat supply

without heat release, the state of the system changes: Although it could be guaranteed by the first
law of thermodynamics that the internal energy, i.e. the temperature, remains constant inside if all
13.3 Energy Conversion—Why Do We Need Entropy? 261

flux of entropy is transferred14 into the system by Q̇ in and a flux of entropy leaves the
system by Q̇ out . To determine the fluxes of entropy Ṡa,in and Ṡa,out , the temperatures
within the system at which the heat exceeds the system boundary are required. As
already mentioned, the machine is operated between a hot and a cold container. The
heat flux Q̇ in enters the system at the hot side, where the temperature is Tmax , Q̇ out
leaves at the cold side, where Tmin is measured, see Fig. 13.4. Since the focus is on
the best case, there is no dissipation and thus no flux of generation of entropy Ṡi
within the system. Hence, the entropy balance in steady state obeys
 
δ Q̇ in Q̇ in Q̇ out δ Q̇ out
Ṡa,in = = = = = Ṡa,out . (13.22)
Tmax Tmax Tmin Tmin

In any case, this equation shows that cooling is necessary to release entropy. If there
were no cooling, the entropy would permanently increase because the heat flux Q̇ in
carries entropy into the system. If entropy were also created inside the machine
through dissipation, even more entropy would have to leave the machine in order
to keep the state value of entropy inside constant. Cooling would therefore have to
be increased even further. Back to the best case: The entropy balance provides the
necessary second equation. It indicates how much cooling is required for steady-state
operation, i.e.
Tmin
Q̇ out = Q̇ in . (13.23)
Tmax

The system of equations is now determined: With two equations, namely Eqs. 13.19
and 13.23, for two unknowns, namely P and Q̇ out , the cycle is completely described.
The maximum power that a thermal power plant releases in the best case is therefore


Tmin Tmin
P = Q̇ in − Q̇ in = Q̇ in 1 − . (13.24)
Tmax Tmax

Obviously, the higher the maximum temperature in the combustion chamber, the more
work can be gained from the given thermal energy flux Q̇ in . The thermal efficiency
of the thermal power plant ηth is defined by the ratio of benefit and effort, i.e.

Benefit P Tmin
ηth = = =1− (13.25)
Effort Q̇ in Tmax

Theorem 13.5 This example shows that the first law of thermodynamics on its own
cannot describe the limits of energy conversion. Only by applying an entropy balance
the thermodynamics of a thermal engine can be fully understood.

the heat supplied is released as work, the entropy, e.g. for ideal gases, does not only depend on the
temperature but also on the specific volume. If the temperature remains constant and the entropy
increases, the specific volume would have to increase constantly, cf. Eq. 13.15.
14 Note that heat is a carrier for entropy.
262 13 Meaning and Handling of Entropy

13.4 The T, s-Diagram

This section is about a new state diagram that is needed to analyse thermodynamic
changes of state in their entirety. In some of the previous examples, the T, s-diagram
has already been addressed, but it is formally introduced in this chapter. It is shown
that with this new diagram it is possible to visualise all process values that have been
introduced so far: Volume/pressure work can be represented in a p, v-diagram, while
heat and dissipation can be visualised in a T, s-diagram.

13.4.1 Benefit of a New State Diagram

At least for single-component/single-phase fluids, Gibb’s phase rule has shown that
two independent state values define a state, cf. Sect. 5.1. Since the specific entropy s
is a state value, the state of a thermodynamic system can be uniquely specified by the
system temperature T and the specific entropy s. Thus, states and changes of state
can be visualised in a T, s-diagram, which contains additional information compared
to a p, v-diagram. In Fig. 13.5 the principle of a T, s-diagram is shown. States (1)
and (2) represent thermodynamic states. These states are explicitly given and it is
possible to derive all the other state values, e.g. pressure p, temperature T , specific
volume v, specific enthalpy h and specific internal energy u, by equations of state. If
the extensive state values are of interest, such as enthalpy H , additional information
regarding the size of the system is required. In addition to the states of a system, it is
also important to understand which path the system takes from one state to another.
Figure 13.5 shows two possible directions the system may take. Let is now focus on
path 1. Obviously, the grey coloured area is the integration of T ds from state (1)
to state (2). In the previous sections, the specific entropy s has been derived as

δq δψ
ds = + . (13.26)
T T

Fig. 13.5 T, s-diagram:


Principle
13.4 The T, s-Diagram 263

Multiplying with T :
T ds = δq + δψ. (13.27)

Integration from state (1) to state (2) leads to the area beneath the change of state in
a T, s-diagram
2 2 2
T ds = δq + δψ. (13.28)
1 1 1

The area under the change of state obviously represents the sum of specific heat q12
and specific dissipation ψ12 , i.e.

2
T ds = q12 + ψ12 (13.29)
1

Since no restrictions have been made regarding the type of fluid, this conclusion
applies equally to ideal gases and real fluids. However, the T, s-diagram proves, that
specific heat as well as specific dissipation are process values, since they are both
path-dependent. For example, if path 2 had been taken, the amount of q12 + ψ12
would be different from path 1 as Fig. 13.5 indicates. Nevertheless, state (2) can be
reached as well.
Figure 13.6 shows the reversible and the irreversible case:
• The reversible case shown in Fig. 13.6a shows the heating of a fluid. The only
mechanism for changing entropy is heat transfer, as there is no entropy generation
due to irreversibilities. Reversible changes of state are always free of dissipation,
see Theorem 7.9 and Example 14.1. Since the specific entropy rises, i.e. s2 > s1 ,
entropy needs to be supplied to the system. This can only be achieved by supply
of heat. In contrast, reversible cooling of the system would lead to s2 < s1 .

Fig. 13.6 T, s-diagram: a Reversible and b irreversible


264 13 Meaning and Handling of Entropy

Fig. 13.7 T, s-diagram: Isochore and isobar

• Due to irreversibility caused by dissipation, entropy is generated in the system,


so that the amount of entropy increases. Heat crossing the system boundary is a
carrier for entropy as well. Depending on the sign of the heat energy, the entropy of
the system can increase or decrease. With regard to Fig. 13.6b, heating or cooling
is possible. In the case of cooling, however, the entropy released must be smaller
than the entropy generated, since s2 > s1 .

13.4.2 Physical Laws in a T, s-Diagram for Ideal Gases

In order to illustrate changes of state it is required to understand how to navigate in


a T, s-diagram. This is clarified in this section. The focus is on the isolines, such
as isobar and isochor. After the specific entropy has been derived in Sect. 12.2, a
combination of the first and second law of thermodynamics leads to Eq. 12.43, which
is also known as the fundamental equation of thermodynamics

T ds = du + p dv (13.30)

respectively
T ds = dh − v d p (13.31)

These equations15 can be applied to derive isobars as well as isochores in a


T, s-diagram, see Fig. 13.7.

15 The fundamental equations of thermodynamics apply to all fluids, i.e. not only to ideal gases.
13.4 The T, s-Diagram 265

Isobaric change of state

For an isobaric change of state, the pressure remains constant so that

d p = 0. (13.32)

Equation 13.31 then reads as


T ds = dh. (13.33)

For ideal gases the caloric equation of state, see Eq. 12.112, can be applied, i.e.

T ds = c p dT. (13.34)

Thus, the slope of the isobar in a T, s-diagram obeys




dT T
= >0 (13.35)
ds p cp

Obviously, the slope is positive and rises with increasing temperature T , cf. Fig. 13.7a.
In order to describe a change of state from (1) to (2), see Fig. 13.7b, Eq. 13.33 needs
to be integrated, so that
2
T ds = h 2 − h 1 . (13.36)
1

With other words, for an isobaric change of state, the area beneath the curve from
(1) to (2) represents the difference of the specific enthalpies h 2 − h 1 , see Fig. 13.7b.

Isochoric change of state

For an isochoric change of state the specific volume remains constant so that

dv = 0. (13.37)

Equation 13.30 then reads as


T ds = du. (13.38)

For ideal gases the caloric equation of state, see Eq. 12.106, can be applied, i.e.

T ds = cv dT. (13.39)

Thus, the slope of the isochore in a T, s-diagram obeys




dT T
= >0 (13.40)
ds v cv
266 13 Meaning and Handling of Entropy

Obviously, the slope is positive and rises with increasing temperature T , see
Fig. 13.7a. Due to
c p = cv + R (13.41)

c p is larger than cv so that




dT dT
> . (13.42)
ds v ds p

Theorem 13.6 For ideal gases, isochores have a larger gradient than isobars in a
T, s-diagram , see Fig. 13.7a.
The gradients cTp and cTv of an isobar and an isochore can even be constructed geo-
metrically for state (1), see Fig. 13.7a. To describe a change of state from (1) to (2 ),
cf. Fig. 13.7b, Eq. 13.38 is integrated so that

2
T ds = u 2 − u 1 . (13.43)
1

With other words, for an isochoric change of state, the area beneath a curve from
(1) to (2 ) represents the difference of the specific internal energies u 2 − u 1 , see
Fig. 13.7b.
Figure 13.8 gives an overview of the isolines in a T, s-diagram. It is already known
that the gradient of an isochore is greater than the gradient of an isobar. Furthermore,
according to states (1) and (2), as given in Fig. 13.8, it is

T2 p2
s2 − s1 = 0 = c p ln − R ln (13.44)
T1 p1

Fig. 13.8 Isobar and


isochore in a T, s-diagram
(ideal gas)
13.4 The T, s-Diagram 267

respectively
T2 p2
c p ln = R ln . (13.45)
T1 p1

Therefore, the pressure in state (2) needs to be greater than in state (1). On the other
hand it is
T2 v2
s2 − s1 = 0 = cv ln + R ln (13.46)
T1 v1

respectively
T2 v2 v1
cv ln = −R ln = R ln . (13.47)
T1 v1 v2

Consequently, the specific volume in state (2) needs to be smaller than in state (1).
The conclusion is:
Theorem 13.7 The higher the isobar in a T, s-diagram, the greater the pressure.
The higher the isochore in a T, s-diagram, the smaller the specific volume.

13.5 Adiabatic, Reversible Change of State

At this point a new change of state is introduced, see Fig. 13.9: The system is supposed
to be adiabatic and free of dissipation, i.e. reversible. Hence, it is

δq = 0 ⇒ δsa = 0 (13.48)

and
δψ = 0 ⇒ δsi = 0. (13.49)

Following Eq. 12.44 it is

δq δψ
ds = + = δsa + δsi = 0 (13.50)
T T

Fig. 13.9 Adiabatic,


reversible change of state
268 13 Meaning and Handling of Entropy

This equation shows that for any adiabatic, reversible change of state, the entropy in
the system remains constant,16 i.e.

s2 = s1 . (13.51)

Such a change of state is denoted isentropic, i.e. ds = 0. Under these conditions the
fundamental equations, see Eqs. 13.30 and 13.31 simplify to

0 = du + p dv (13.52)

and
0 = dh − v d p. (13.53)

By rearranging and combining these two equations one obtains

du p dv
=− (13.54)
dh v dp

This equation can be applied to any type of fluid, as no restrictions have been made
so far. However, the focus is now on ideal gases: The caloric equations for u and h
for an ideal gas are, see Sect. 12.3:


∂p
du = cv (v, T ) · dT + T − p · dv = cv dT (13.55)
∂T v

and

∂v
dh = c p ( p, T ) · dT + −T + v · d p = c p dT. (13.56)
∂T p

cp
Hence, Eq. 13.54 simplifies for ideal gases with κ = cv
, i.e.

1 p dv
=− . (13.57)
κ v dp

Separation of the variables leads to

dp dv
− =κ . (13.58)
p v

16No entropy is transported across the system boundary by heat and there is no entropy generation
within the system. Similar to a bathtub, which has no inflow or outflow, the entropy in the system
must therefore be constant.
13.5 Adiabatic, Reversible Change of State 269

An integration results in

− ln p2 + ln p1 = κ (ln v2 − ln v1 ) (13.59)

respectively
p1 v2
ln = κ ln . (13.60)
p2 v1

The reformulation is as follows



κ
p1 v2
= . (13.61)
p2 v1

In other words, an adiabatic, reversible, i.e. isentropic, change of state for ideal gases
obeys
pvκ = const. (13.62)

In a different notation it is
p1 v1κ = p2 v2κ . (13.63)

Applying the thermal equation of state leads to


κ−1
κ
T2 p2
= (13.64)
T1 p1

As described before, an isentropic change of state is defined as ds = 0. However,


there are two options for achieving an isentropic change of state:
• The change of state is adiabatic and reversible, e.g. according to Fig. 13.9:

δq δψ
ds = + = 0. (13.65)
T T
 
=0 =0

• If dissipation occurs, i.e. the change of state is irreversible, the change of state can
still be isentropic. According to Eq. 12.44 cooling is required:

δq δψ
ds = + = 0. (13.66)
T
 T

<0 >0

Hence, it follows
q12 = −ψ12 . (13.67)

Under these conditions, the heat released, i.e. the cooling, must be equal to the
dissipated energy. The entropy in the system would then remain constant.
270 13 Meaning and Handling of Entropy

A heated system can never be isentropic, because under these conditions entropy in
the system would have to be destroyed. This is not possible, since si,12 can only be
positive or zero.
Problem 13.1 An ideal gas is compressed reversibly in a horizontal cylinder. Two
different options shall be investigated:
• (1) → (2): Isothermal compression
• (1) → (3): Adiabatic compression, so that v3 = v2
Sketch the two changes of state in a p, v-diagram and in a T, s-diagram!

Solution

Let us start with the first law of thermodynamics for the isothermal case:

Q 12 + W12 = U2 − U1 = mcv (T2 − T1 ) = 0. (13.68)

Hence, it is
Q 12 = −W12 . (13.69)

The partial energy equation reads as

2
W12 = WV,12 = − p dV > 0. (13.70)
1

The work W12 is positive, since for a compression it is dV < 0. This leads to

Q 12 = −W12 < 0. (13.71)

For an isothermal compression, heat needs to be released. In a p, v-diagram, the


change of state follows an isotherm, see Fig. 13.10. The volume decreases and accord-
ingly, since the mass in a closed system is constant, the specific volume also decreases.
Thus, the new state (2) is to the left of state (1). The mathematical function p = f (v)
for an isotherm can be derived from the thermal equation of state

pv = RT = const. (13.72)

and reads as
RT p1 v1 1
p (v) = = ∝ . (13.73)
v v v
In a T, s-diagram, an isothermal change of state must follow a horizontal line. Since
the system needs to be cooled, entropy is released by the system, so that the specific
entropy s2 is smaller than the initial specific entropy s1 . Therefore, at the same
temperature, state (2) lies to the left of state (1), see Fig. 13.10. A comparison of the
13.5 Adiabatic, Reversible Change of State 271

Fig. 13.10 Reversible compression of an ideal gas: Isothermal versus adiabatic according to Prob-
lem 13.2

isobars p1 and p2 in the T, s-diagram shows that the location of state (2) makes sense,
since the isobar p2 lies above the isobar p1 . This correlates with the p, v-diagram as
well. The specific heat released is represented by the hatched area under the curve
(1) → (2) in the T, s-diagram.
For an adiabatic, reversible, i.e. isentropic, change of state (1) → (3) the first law
of thermodynamics obeys

Q 13 +W13 = U3 − U1 = mcv (T3 − T1 ) . (13.74)



=0

The work follows the partial energy equation

3
W13 = WV,13 = − p dV > 0. (13.75)
1

Hence, the first law of thermodynamics states

W13 = U3 − U1 = mcv (T3 − T1 ) > 0. (13.76)

Consequently, for an adiabatic, reversible change of state the temperature rises:

T3 > T1 . (13.77)

The function for an isentropic change of state has been recently derived and results in

pvκ = const. (13.78)


272 13 Meaning and Handling of Entropy

In other words, the function p (v) is

p1 v1κ 1
p (v) = ∝ κ. (13.79)
vκ v
Apparently, due to κ > 1, the pressure increases faster with decreasing specific
volume in an isentropic change of state than in an isothermal change of state, cf.
Fig. 13.10. This corresponds to the considerations of the first law of thermodynam-
ics. Note that the thermal equation of state can be applied over the entire curve from
(1) → (3), in case the change of state is quasi-static17 :

pv = RT = const. (13.80)

Using Eq. 13.80, the temperature of the system at any point along the curve can be
calculated using Eq. 13.79. However, in a T, s-diagram state (1) → (3) follows a
vertical line, since the entropy stays constant. Due to T3 > T1 , state (3) is above state
(1). This also correlates with the isobars p1 and p3 in the T, s-diagram.

13.6 Polytropic Change of State

In Problem 13.2, two important changes of state, namely an isothermal and an isen-
tropic change of state, have been compared. It has been shown that an isentropic
change of state can be realised with two alternatives: A system with dissipation can
only be isentropic if heat is released. Furthermore, an adiabatic system can be isen-
tropic if it is free of dissipation. Heated systems can never be isentropic. Figure 13.10
shows both an isothermal and an isentropic change of state in a T, s-diagram and in
a p, v-diagram accordingly.
However, these two cases are special cases in technical applications: As everyday
experience shows, there are no perfectly insulated, i.e. adiabatic systems, since heat
exchange with the environment18 can never be completely avoided.
Let us assume, that T1 is identical with ambient temperature, see Fig. 13.11. In
case (1) → (4) the thermal insulation of the system is not perfect. However, the
temperature of the system increases during compression. Due to the thermal coupling
with the environment, heat is released to the cooler environment. Consequently, the
temperature in the final state T4 is lower than the temperature T3 resulting from
adiabatic compression. If the system is adiabatic but irreversible, the final temperature
is even greater than T3 . This case is discussed as change of state (1) → (5) in
Fig. 13.14. These changes of state are defined as polytropic.
As these examples show, different temperatures can be achieved through compres-
sion, it just depends on the process control. In order to calculate the process values

17 The prerequisite for a reversible change of state is that it is also quasi-static.


18 In case the system has a different temperature than environment.
13.6 Polytropic Change of State 273

Fig. 13.11 Polytropic


change of state for an ideal
gas in a p, v-diagram

volume work or technical work, a mathematical function p = f (v) is required. Two


cases have already been discussed:
• Isothermal change of state (ideal gas)

pv = RT = const. (13.81)

The explicit notation of the required function is

const.
p (v) = . (13.82)
v
In thermodynamics, however, an implicit notation is usually used, i.e.

pv = const. (13.83)

• Isentropic change of state (ideal gas)

This case has already been discussed in Sect. 13.5. Its implicit notation obeys

pvκ = const. (13.84)

Isentropic changes of state are important thermodynamic benchmarks. In Chap. 21,


for example, in which flow processes such as supersonic flows are studied, they
play an important role: In case the flow is very fast, there is almost no heat transfer
to the environment, i.e. the flow is nearly adiabatic. Consequently, neglecting the
dissipation, the flow is approximately isentropic.

• Polytropic change of state (ideal gas)

Obviously, a polytropic change of state (1) → (4) in a p, v-diagram has a different


gradient than an isothermal/isentropic change of state, see Fig. 13.11. Comparing
274 13 Meaning and Handling of Entropy

Table 13.1 Polytropic change of state—Technical meaning for ideal gases


Polytropic exponent n Change of state
0 pv 0 = const. ⇒ p = const.
1 pv 1 = const. ⇒ T = const.
κ pv κ = const. ⇒ isentrop
∞ pv ∞ = const. ⇒ p 1/∞ v 1 = const. ⇒ v = const.

Fig. 13.12 Polytropic change of state for an ideal gas in a p, v- and in a T, s-diagram according
to [1]

the correlations for an isotherm, see Eq. 13.83, and an isentrope, see Eq. 13.84, the
gradient is obviously given by the exponent with respect to the specific volume v.
Thus, a polytrope is defined by:

pvn = const. (13.85)

Depending on the process n can be any number −∞ < n < ∞. However, some
cases have a special technical meaning, see Table 13.1.
In Fig. 13.12 these cases are shown in both a p, v- and a T, s-diagram. Thus, e.g.
for case (1) → (4), see Fig. 13.11, it is

1 < n < κ, (13.86)

since the polytrope lies between the isotherm and the isentrope. However, addi-
tionally the thermal equation of state

pv = RT (13.87)

can be applied for all cases, if the gas is ideal and in thermodynamic equilibrium.19
This leads to a different notation of the polytropic change of state:

19 Note that changes of state are assumed to be quasi-static.


13.6 Polytropic Change of State 275

 n−1   n−1
T1 v2 p1 n
= = (13.88)
T2 v1 p2

In technical processes it is assumed, that n = const. for a change of state,20 see [11].
Depending on the technical problem, it makes sense to calculate the polytropic
exponent n in order to determine the process values work or technical work. If the
states (1) and (2) are thermodynamically unambiguous, the exponent can be easily
derived from Eq. 13.85, i.e.
ln pp2
n = v11 . (13.89)
ln v2

Once, the direction of the change of state is determined by the polytropic exponent
n, the following process values can be calculated:

– Specific volume work (closed system)

2 2
n−1 
dv p1 v1 v1
w12,V = − p dv = − p1 v1n = −1 (13.90)
vn n−1 v2
1 1

– Specific pressure work (open system)

2 2
n−1 
1/n −1/n p1 v1 v1
y12 = v dp = v1 p1 p dp = n −1 (13.91)
n−1 v2
1 1

• Polytropic change of state (real fluids)

Although real fluids are not treated until Part II, the definition of a polytropic
change of state for real fluids follows according to Eq. 13.85, i.e.

pvn = const. (13.92)

However, for real fluids, the thermal equation of state pv = RT is not fulfilled.
Consequently, Eq. 13.88 can not be applied for real fluids. Furthermore, an isother-
mal change of state does not obey n = 1 and an isentropic change of state does
not follow n = κ. Consequently, Table 13.1 only counts for ideal gases.
An overview of possible changes of state for ideal gases is given in Fig. 13.13.

20However, if the external influence varies over time, n may also vary, see for example Problem
13.13.
276 13 Meaning and Handling of Entropy

Fig. 13.13 Overview ideal gases

Problem 13.2 An ideal gas is compressed in a horizontal cylinder. Several changes


of state are to be investigated:
• (1) → (2): isothermal (reversible)
• (1) → (2 ): isothermal (irreversible)
• (1) → (3): adiabatic (reversible)
• (1) → (4): polytropic (reversible), 1 < n < κ
• (1) → (5): adiabatic (irreversible)
The final volume shall be the same in all cases. Please illustrate all changes of state
in a p, v- and T, s-diagram. There shall be no friction for the piston. Irreversibility
is due to dissipation within the gas.
Solution

The diagrams are shown in Fig. 13.14.


13.6 Polytropic Change of State 277

Fig. 13.14 Polytropic change of state for an ideal gas - Compression (closed system), see Problem
13.2

Problem 13.3 An ideal gas expands in a horizontal cylinder. Several changes of


state are to be investigated:
• (1) → (2): isothermal (reversible)
• (1) → (2 ): isothermal (irreversible)
• (1) → (3): adiabatic (reversible)
• (1) → (4): adiabatic (irreversible)
The final pressure shall be the same in all cases, since the system strives for mechan-
ical equilibrium with the environment. Please illustrate all changes of state in a p, v-
and T, s-diagram. There shall be no friction for the piston. Irreversibility is due to
dissipation within the gas.

Solution

The diagrams are shown in Fig. 13.15. Anyhow, let us have a closer look at state
(4), that is achieved by an adiabatic but irreversible change of state. Since, the gas
inside the horizontal cylinder expands, the piston compresses the environment. It is
obvious that volume work is released into the environment. This work is provided
by the expanding gas, so that the work W14 must be negative.21 Hence, the partial
energy equation is

21 W14 is identical with the work supplied to the environment, since the piston moves horizontally,
i.e. W14 is not utilised to lift a mass in a gravitational field. There is also no friction between the
piston and the cylinder wall.
278 13 Meaning and Handling of Entropy

Fig. 13.15 Polytropic change of state for an ideal gas—Expansion (closed system), see Problem
13.3

4
W14 = W14,V + 14 = − p dV + 14 < 0. (13.93)
1

The first law of thermodynamics follows

W14 + Q 14 = U4 − U1 = mcv (T4 − T1 ) (13.94)


 
<0 =0

so that T4 < T1 . However, according to the T, s-diagram in Fig. 13.15, temperature


T4 must be greater than T3 , which is reached by an isentropic change of state. This
is due to p3 = p4 and s4 > s3 . The specific entropy rises because of the specific
dissipation ψ14 .

Problem 13.4 A cooled compressor sucks in air out of the environment ( penv =
3
1 bar, Tenv = 288 K) with a volume flow rate of 3500 mh . As the air flows through
the compressor, it is compressed in a stationary process to a final pressure of 3.5 bar.
As a result, the temperature of the air in the outlet tube (d = 160 mm) reaches 393 K.
During this process a heat flux of 49 kW is released.
(a) Calculate the power consumption of the compressor.
(b) Calculate the power consumption of a compressor at the same pressure ratio
and the same air mass flow for the case that the compression is isentropic and
frictionless.
13.6 Polytropic Change of State 279

Fig. 13.16 Sketch of the


compressor, see Problem
13.4

The isentropic exponent is κ = const. = 1.4 and the air can be regarded as an ideal
gas (R = 287 kgJK ).

Solution

A compressor is a component applied to increase the pressure of a gas. To achieve


this, work must be supplied, see Fig. 13.16.
(a) The first law of thermodynamics is used to calculate the power consumption.
According to the sketch it reads



1 1
Q̇ 12 + Pt,12 + ṁ h 1 + c12 + gz 1 = ṁ h 2 + c22 + gz 2 . (13.95)
2 2

The requested power consumption is


 
1 2 
Pt,12 = ṁ h 2 − h 1 + c − c1 + g (z 2 − z 1 ) − Q̇ 12 .
2
(13.96)
2 2

The mass flow rate can be calculated with the thermal equation of state for state
(1), since the volume flow rate22 V̇1 is given, i.e.

m 1 RT1 = p1 V1 . (13.97)

The time derivative results in

ṁ 1 RT1 = p1 V̇1 . (13.98)

Hence, the mass flow rate in steady state is

p1 V̇1 kg
ṁ = ṁ 1 = = 1.176 . (13.99)
RT1 s

22 Note that V̇1 = V̇2 . Steady state means that the mass flow is constant. Due to the compressibility
of air, the volume flow is not constant.
280 13 Meaning and Handling of Entropy

Since state (1) is ambient state, its velocity23 is c1 = 0. The velocity c2 can be
calculated by applying the equation of continuity:

c2 A2 c2 π d22 p2 c2 π d22
ṁ = ρ2 c2 A2 = = = . (13.100)
v2 4v2 4RT2

The velocity c2 is
4ṁ RT2 m
c2 = = 18.84 . (13.101)
p2 π d2
2 s

The specific heat capacity is:

κ J
cp = R = 1004.5 . (13.102)
κ −1 kg K

Finally, the power consumption can be calculated by applying the caloric equa-
tion of state24 :
 
1 2 
Pt,12 = ṁ c p (T2 − T1 ) + c2 − c12 − Q̇ 12 = 173.2 kW (13.103)
2

(b) The process shall be isentropic (ds = 0) and reversible (δsi = 0). Thus, it
follows:
ds = δsi + δsa ⇒ δsa = ds − δsi = 0. (13.104)

Consequently, the change of state needs to be adiabatic, so that Q̇ 12 = 0. The


first law of thermodynamics for the change of state (1) → (2 ) obeys
⎡ ⎤

1 2 
Pt,12 = ṁ ⎣c p (T2 − T1 ) + c  − c12 + g (z 2 − z 1 )⎦ . (13.105)
2 2   
=0

In order to determine the final temperature T2 for an isentropic change of state
Eq. 13.64 can be applied. As the pressure ratio is the same as in (a), the pressure
p2 is equal to p2 , i.e.

κ−1
κ
p2
T = T1
2 = 412 K. (13.106)
p1

Since the mass flow rate shall be the same as in (a), the velocity c2 is

23 But even if the velocity is almost zero, there can still be a airflow according to V̇1 = c1 A1 if the
cross-section A1 → ∞, i.e. the environment is considered to be very large.
24 Since there is no information about the vertical position of the inlet and outlet, the change in

potential energy is ignored.


13.6 Polytropic Change of State 281

4ṁ RT2 m
c2 = = 19.7 . (13.107)
p2 π d2
2 s

Finally, the power consumption in case (b) is smaller than in case (a). It follows:
 
1 2 
Pt,12 = ṁ c p (T2 − T1 ) + c  − c1 = 146.4 kW
2
(13.108)
2 2

13.7 Entropy Balancing

To evaluate thermodynamic processes, it is important to balance not only energy


but also entropy. The entropy balance is key to understanding the direction that
energy conversion can take and its restrictions. Therefore, the focus is first on closed
systems, followed by open systems. However, thermodynamic cycles are treated
separately in Chap. 15, as they are part of the classical formulation of the second law
of thermodynamics.

13.7.1 Entropy Balance for Closed Systems

The analogy between energy balancing and entropy balancing has been elaborated
in Sect. 13.2. With Figs. 13.2 and 13.3 the principle of entropy balancing in a closed
system has been introduced in comparison to the first law of thermodynamics, see
also Fig. 13.17. In the section following, the most important findings are summarised
and discussed with a simple example, cf. Example 13.1. It has been shown that heat
passing the system boundary is a carrier for entropy, i.e.

Fig. 13.17 Energy and entropy balance for a closed system


282 13 Meaning and Handling of Entropy

2 2
δQ δQ
δSa = ⇒ δSa = Sa,12 = . (13.109)
T T
1 1

The generation of entropy within the system is due to dissipation and obeys

2 2
δ δ
δSi = ⇒ δSi = Si,12 = . (13.110)
T T
1 1

In these equations, T is the temperature within the system where the heat passes the
boundary respectively where the dissipation takes place. In a homogeneous system,
the heat passing the boundary and the dissipation are at the same uniform temperature.
The change in the state value entropy of the system follows

dS = δSi + δSa (13.111)

The amount of entropy is influenced by both the entropy that crosses the system
boundary and the entropy that is generated within the system. The change in entropy
within the system is therefore determined by integration:

S2 − S1 = m(s2 − s1 ) = Si,12 + Sa,12 (13.112)

Compare with the analogous notation of the first law of thermodynamics25 :

U2 − U1 = m(u 2 − u 1 ) = Q 12 + W12 (13.113)

For both equations, i.e. energy and entropy balance, the caloric equations of state can
be applied. In the case of an ideal gas, the caloric equations of state are as follows

T2 v2 T2 p2
s2 − s1 = cv · ln + R · ln = c p · ln − R · ln (13.114)
T1 v1 T1 p1

respectively
u 2 − u 1 = cv · (T2 − T1 ) . (13.115)

Example 13.1 (Isothermal irreversible compression) This example answers the


question by what amount the entropy of an ideal gas changes during an isother-
mal, irreversible compression, cf. Fig. 13.18. The thermal equation of state for an
isothermal change of state obeys

25 If the change in outer energies is neglected, e.g. because a horizontal cylinder is at rest in state
(1) and in state (2).
13.7 Entropy Balancing 283

Fig. 13.18 Isothermal


irreversible compression of
an ideal gas

pv = RT = const. (13.116)

Thus, it is
p1 v1 = p2 v2 . (13.117)

Since for a compression it is v2 < v1 , it follows

p2 > p1 . (13.118)

The first law of thermodynamics for that case yields

W12 + Q 12 = U2 − U1 + E a,12 . (13.119)

• The partial energy equation reads as

W12 = WV,12 + Wmech,12 +12 . (13.120)


  
=E a,12

Mechanical work Wmech,12 is required to change the outer energy by E a,12 .

• The caloric equation of state is

U2 − U1 = mcv (T2 − T1 ) = 0. (13.121)

Hence, the first law of thermodynamics follows

12 + Q 12 = −WV,12 . (13.122)

The volume work WV,12 can be calculated according to Fig. 13.13 so that

p2
12 + Q 12 = −WV,12 = −m RT ln . (13.123)
p1

Let us now focus on the change of entropy in this change of state. It follows
284 13 Meaning and Handling of Entropy

S − S = Si,12 + Sa,12 . (13.124)


 2  1   
Alternative 1 Alternative 2

There are two alternatives, to calculate the change of entropy:


• Alternative 1 follows the caloric equation of state, i.e. it investigates states (1) and
(2), i.e.
 
T2 p2
S2 − S1 = m (s2 − s1 ) = m c p · ln − R · ln . (13.125)
T1 p1

Since the change of state is isothermal, temperature T2 and temperature T2 are


equal. Thus, it is
p2
S2 − S1 = −m R · ln (13.126)
p1

• Alternative 2 investigates the cause for the change of entropy, i.e.

2 2
δ δQ
S2 − S1 = Si,12 + Sa,12 = + . (13.127)
T T
1 1

Since the temperature T is constant, the equation simplifies to

2 2
1 1
S2 − S1 = Si,12 + Sa,12 = δ + δQ (13.128)
T T
1 1

respectively
12 Q 12
S2 − S1 = Si,12 + Sa,12 = + . (13.129)
T T
The combination with Eq. 13.122 leads to

p2
S2 − S1 = Si,12 + Sa,12 = −m R · ln (13.130)
p1

However, both alternatives lead to the same result:

p2
S2 − S1 = −m R ln <0 (13.131)
p1

The change of state is illustrated as change of state (1) → (2 ) in Fig. 13.14.
13.7 Entropy Balancing 285

13.7.2 Entropy Balance for Open Systems

The law of conservation of energy for open systems has already been derived in
Sect. 11.3.1, for both transient and steady states. With the energy fluxes that pass the
system boundary, it is possible to balance the amount of energy within the system.
This principle is illustrated in Fig. 13.19a. In this sketch, however, the kinetic and
potential energy at the inlet and outlet are ignored, only the enthalpies are shown. It
applies that energy fluxes entering the system increase the amount of energy in the
system, while energy fluxes leaving the system decrease its energy content. In fact,
this principle26 can be applied to any extensive state value Z and has been discussed
recently in Sect. 3.2.5.
Consequently, the entropy follows accordingly, see Fig. 13.19b: Any entropy flux
that crosses the system boundary and enters the system causes the entropy in the
system to increase over time. Entropy fluxes leaving the system cause the entropy
inside to decrease over time. Furthermore, in contrast to the energy balance, there
can be an additional source term for the entropy, i.e. an entropy generation rate Ṡi due
to a flux of dissipated energy. This source term also causes the entropy in the system
to increase with time. Dissipation  leads to generation of entropy Si ; a flux of dis-
sipation ˙ in open systems causes a flux of entropy generation Ṡi , see Sect. 12.2, i.e.

˙ → Ṡi .
 → Si respect.  (13.132)

Let us take a closer look at the entropy fluxes that cross the system boundary. Heat as
an entropy carrier causes entropy Sa to flow into the system when the system is heated
or to flow out of the system when it is cooled. In case of a heat flux, e.g. in an open sys-
tem that is operated permanently, an entropy flux Ṡa is transported, see Sect. 12.2, i.e.

Q → Sa respect. Q̇ → Ṡa . (13.133)

Fig. 13.19 a Energy and b entropy balance—open system

26 Bathtub principle.
286 13 Meaning and Handling of Entropy

In contrast to closed systems, due to the mass flux entropy is also carried convectively
into or out of the system: Each mass flow carries state values.27 However, extensive
state values28 which are thus supplied to the system increase the state value inside.
In contrast, extensive state values which are released due to the mass flux decrease
the state value within the system: Consequently, in addition to Ṡi and Ṡa , the entropy
of the system is increased with the inflow Ṡ1 = ṁ 1 s1 and decreased with the outflow
Ṡ2 = ṁ 2 s2 . Therefore, according to Sect. 11.3.1, the conclusion for the temporal
change of entropy in the system, cf. Fig. 13.19b, is29

dS
= ṁ 1 s1 + Ṡa + Ṡi − ṁ 2 s2 (13.134)
dt    
IN OUT

The term on the left side of the Eq. 13.134 takes into account the temporal change
of entropy within the system. This change of the entropy of the system is caused by
the fluxes respectively the source term, which are summarised on the right hand side
of Eq. 13.134. In case the fluxes in equal the fluxes out there is no temporal change
inside the system.30

Steady state flow systems

Let us assume that the system shown in Fig. 13.19 is in steady state. In this case the
continuity equation obeys
ṁ 1 = ṁ 2 . (13.135)

With other words, the mass flux into the system needs to be equal to the mass flux
leaving the system. Regarding the entropy balance, cf. Eq. 13.134, for steady state
conditions dS
dt
needs to be zero, i.e.

0 = ṁ 1 s1 + Ṡa + Ṡi − ṁ 2 s2 . (13.136)

This is due to the definition of steady state, i.e. all state values within the system must
be constant in time, even though they may vary locally. However, this does not only
affect the state values pressure p, temperature T , but also the state value entropy S.
Consequently, the entropy balance can be represented in the following notation

ṁ 1 s1 + Ṡa + Ṡi = ṁ 2 s2 . (13.137)


   
IN OUT

27 As already discussed, two independent state values represent the state unequivocally. At least for
single component fluids without phase change, see Sect. 5.1.
28 Extensive state values are countable, i.e. they can be represented by bubbles.
29 Bathtub principle.
30 According to the arrows in Fig. 13.19.
13.7 Entropy Balancing 287

In other words:
Entropy flux in = Entropy flux out

Note that the source term Ṡi is treated as a flow into the system as it increases the
entropy of the system.
Multiple inlets/outlets
Open systems are of course not limited to one outlet or inlet. In many technical
applications, mass flows are supplied at several inlets, or a mass flow is withdrawn,
so that several inlets/outlets must also be considered. The general entropy balance
for such a system results from our considerations in Sect. 13.7.2:

dS
= ṁ i si + Ṡa + Ṡi − ṁ j s j (13.138)
dt in,i out, j

In steady state this equation simplifies to



0= ṁ i si + Ṡa + Ṡi − ṁ j s j (13.139)
in,i out, j

respectively
ṁ i si + Ṡa + Ṡi = ṁ j s j . (13.140)
in,i out, j

As before, the conclusion is for multiple inlet and outlet systems in steady state:
Entropy flux in = Entropy flux out

Example 13.2 This example deals with an open system in steady state shown in
Fig. 13.20.
• First, it is advisable to start with the equation of continuity:

Mass flux in = Mass flux out

Fig. 13.20 Entropy balance—steady state flow system, multiple inlets/outlets


288 13 Meaning and Handling of Entropy

In this case:
ṁ 1,a + ṁ 1,b = ṁ 2 . (13.141)

• However, the next step is the first law of thermodynamics:

Energy flux in = Energy flux out


Applied to Fig. 13.20:



1 2 1 2
Q̇ + Pt + ṁ 1,a h 1,a + c1,a + gz 1,a + ṁ 1,b h 1,b + c1,b + gz 1,b =
2 2


(13.142)
1
ṁ 2 h 2 + c22 + gz 2 .
2

Let us disregard the kinetic and potential energy. By transforming and combining
with the continuity equation, i.e. Eq. 13.141, we get
   
Q̇ + Pt = ṁ 1,a h 2 − h 1,a + ṁ 1,b h 2 − h 1,b . (13.143)

This notation is advantageous because the caloric equation of state can be applied
for the differences in the specific enthalpies.31

• Finally, the entropy is balanced, i.e. the second law of thermodynamics is applied:

Energy flux in = Energy flux out

In this example it is

Ṡa + Ṡi + ṁ 1,a s1,a + ṁ 1,b s1,b = ṁ 2 s2 . (13.144)

Applying Eq. 13.141 and rearranging in order to get differences for the specific
entropy s leads to
   
Ṡa + Ṡi = ṁ 1,a s2 − s1,a + ṁ 1,b s2 − s1,b . (13.145)

The caloric equations of state can be applied if it is an ideal gas, see e.g. Eq. 12.54,
i.e.
T2 p2
s2 − s1,a = c p ln − R ln (13.146)
T1,a p1,a

and
T2 p2
s2 − s1,b = c p ln − R ln . (13.147)
T1,b p1,b

31 Note that the caloric equations of state always specify the change of a caloric state value.
13.7 Entropy Balancing 289

13.7.3 Thermodynamic Mean Temperature

It has been shown that heat that crosses the system boundary leads to an entropy
transfer. When a heat flux occurs, an entropy flux passes the system boundary:
Q 12 Q̇
Sa,12 = respect. Ṡa = . (13.148)
T T
In addition, entropy is generated in the system when energy is dissipated. If there is
a flux of dissipated energy, a flux of entropy is generated, i.e.
12 ˙

Si,12 = respect. Ṡi = . (13.149)
T T
In this section, these two mechanisms are examined in more detail. It has already
been mentioned that the temperature T is the temperature within the system at which
heat passes the system boundary respectively at which dissipation occurs. Obviously,
T does not need to be locally constant. Consequently, temperature, heat and dissipa-
tion can depend on location and time. Accordingly, the entropy transferred by heat
and the entropy generated in a system can be functions of place and time. To clarify
how to deal with this phenomenon, the focus is first on stationary open systems and
then on closed systems.

Steady state flow systems

In open systems the determination of Ṡa and Ṡi sometimes causes difficulties, since
the temperature inside the system is non-uniform, e.g. the inlet temperature T1 can
be different from the outlet temperature T2 , cf. Fig. 13.21. However, the system is
in local thermodynamic equilibrium, see Sect. 4.4. Since the system is supposed
to be in steady state, the temperature profile is constant with time. Accordingly, if
dissipation occurs inside the system, this is at a non-uniform, but temporally constant
temperature.
In order to solve this problem, the system is therefore spatially discretised, see
Fig. 13.21. Transferred heat and dissipation now face a varying temperature Ti from
element to element. Anyhow, the entire heat being transferred is

Fig. 13.21 Thermodynamic mean temperature—steady state flow system


290 13 Meaning and Handling of Entropy

2
Q̇ 12 = δ Q̇ i (13.150)
1

and the entire dissipation is


2
˙ 12 =
 ˙ i.
δ (13.151)
1

Entropy is transferred differentially into32 each element by heat transfer:

δ Q̇ i
δ Ṡa = . (13.152)
Ti

The total entropy that is carried with the heat is

2 2
δ Q̇ i
Ṡa = δ Ṡa = . (13.153)
Ti
1 1

In order to solve this integral a thermodynamic mean temperature Tm is defined, that


is supposed to be constant within the entire system, see Fig. 13.21:

2 2
δ Q̇ i 1 Q̇ 12
Ṡa = ≡ δ Q̇ i = . (13.154)
Ti Tm Tm
1 1

This representative temperature is

Q̇ 12
Tm = (13.155)
Ṡa

The dissipation can be treated accordingly, i.e.

2 2 ˙i
δ
Ṡi = δ Ṡi = . (13.156)
Ti
1 1

With the already introduced thermodynamic mean temperature Tm it results in

32 Or out of the element, in case the system releases heat.


13.7 Entropy Balancing 291

2 ˙i 2 ˙ 12
δ 1 
Ṡi = ≡ ˙i =
δ . (13.157)
Ti Tm Tm
1 1

This equation can be solved for Tm as well

˙ 12

Tm = (13.158)
Ṡi

The combination of Eqs. 13.155 and 13.158 results in


 
˙ 12 + Q̇ 12 .
Tm Ṡa + Ṡi =  (13.159)

Hence, the thermodynamic mean temperature also obeys

˙ 12 + Q̇ 12

Tm = (13.160)
Ṡa + Ṡi

According to Eq. 13.140 for an open system with several inlets/outlets in steady state,
the following applies
Ṡa + Ṡi = ṁ j s j − ṁ i si . (13.161)
out, j in,i

Thus, in such a case the thermodynamic mean temperature follows

˙ 12 + Q̇ 12
Tm =   (13.162)
out, j ṁ j s j − in,i ṁ i

This equation can be simplified for systems with a single inlet and single outlet, i.e.

ψ12 + q12
Tm = (13.163)
s2 − s1

Closed systems

The change of state for a closed system (1) → (2) can be initiated by process values
work and heat. On its way, the temperature of the system may change from T1 to T2 as
heat passes the system boundary. Heat is therefore facing a time-varying temperature.
It is obvious that this varying temperature has an impact on the entropy that passes the
system boundary. In order to calculate the entropy carried with the heat, the change
of state is discretised into sub-steps, see Fig. 13.22. The entire heat follows
292 13 Meaning and Handling of Entropy

Fig. 13.22 Thermodynamic


mean temperature—closed
system

2
Q 12 = δ Qi . (13.164)
1

However, the entropy passing in each sub-step is

δ Qi
δSa,i = . (13.165)
Ti

To calculate the total entropy, this equation must be integrated over all substeps, i.e.

2 2
δ Qi
Sa,12 = δSa,i = . (13.166)
Ti
1 1

Introducing a representative, constant thermodynamic mean temperature Tm results in

2 2
δ Qi 1 Q 12
Sa,12 = ≡ δ Qi = . (13.167)
Ti Tm Tm
1 1

Hence, the thermodynamic mean temperature is

Q 12
Tm = (13.168)
Sa,12
13.7 Entropy Balancing 293

Furthermore, during the change of state, dissipation33 may occur accordingly at


varying temperature. Hence, it affects the generated entropy. Analogous to the heat
it is
2
12 = δi . (13.169)
1

However, the entropy generated in each sub-step yields

δi
δSi,i = . (13.170)
Ti

To calculate the total entropy, this equation must be integrated over all substeps

2 2
δi
Si,12 = δSi,i = . (13.171)
Ti
1 1

Introducing a representative, constant thermodynamic mean temperature Tm results in

2 2
δi 1 12
Si,12 = ≡ δi = . (13.172)
Ti Tm Tm
1 1

Hence, the thermodynamic mean temperature is

12
Tm = (13.173)
Si,12

The combination of Eqs. 13.168 and 13.173 leads to


 
Tm Sa,12 + Si,12 = 12 + Q 12 . (13.174)

Thus, the following also applies to the thermodynamic mean temperature

12 + Q 12
Tm = (13.175)
Sa,12 + Si,12

Due to Eq. 13.112 it follows accordingly

Q 12 + 12 q12 + ψ12


Tm = = (13.176)
S2 − S1 s2 − s1

33 Note that dissipation is a process value as well.


294 13 Meaning and Handling of Entropy

13.7.4 Entropy and Process Evaluation

In this section, an evaluation of thermodynamic processes is clarified. The previous


sections have shown that entropy generation and entropy transport by heat exchange
can be characterised as follows:
• Process values Sa respect. Ṡa
Same sign as the heat passing the system boundary:
– δSa = 0 ⇒ Sa,12 = 0 respect. Ṡa = 0 (adiabatic process)
– δSa > 0 ⇒ Sa,12 > 0 respect. Ṡa > 0 (system is heated)
– δSa < 0 ⇒ Sa,12 < 0 respect. Ṡa < 0 (system is cooled)
• Process value Si respect. Ṡi
Quantitative measure of the degree of irreversibility:
– δSi = 0 ⇒ Si,12 = 0 respect. Ṡi = 0 (reversible process)
– δSi > 0 ⇒ Si,12 > 0 respect. Ṡi > 0 (irreversible process)
– δSi < 0 ⇒ Si,12 < 0 respect. Ṡi < 0 (impossible process)
Theorem 13.8 Therefore, the generation of entropy is utilisied for process evalua-
tion.

Example 13.3 Let us take a closer look at the beer crate problem in Fig. 13.23. We
have learned so far that it requires mechanical work to lift the beer in a gravity field
from position z 1 to position z 2 . Everyday experience shows that this change of state
is rather strenuous.
An inventor claims to have a special device that supplies heat to the beer without
changing its internal state, i.e. temperature and pressure of the beer remain con-
stant. His idea is that all the thermal energy is converted into potential energy. The
task is to investigate whether such a device can exist. According to the first law of
thermodynamics for closed systems

1  
W12 +Q 12 = U2 − U1 + m c22 − c12 +mg (z 2 − z 1 ) (13.177)

=0
2  
=0

Fig. 13.23 Lifting a crate of


beer
13.7 Entropy Balancing 295

in combination with the caloric equation of state

Q 12 = mc (T2 − T1 ) +mg (z 2 − z 1 ) (13.178)


  
=0

it is
Q 12 = mg (z 2 − z 1 ) . (13.179)

According to the first law of thermodynamics, it would therefore be possible to


convert all the thermal energy into potential energy. Nevertheless, let us apply the
partial energy equation as well:

W12 = 0 = WV,12 + Wmech,12 +12 . (13.180)


     
=0 mg(z 2 −z 1 )>0

Solving for the dissipation results in

12 = −mg (z 2 − z 1 ) < 0. (13.181)

This is not possible, since dissipation can never be negative, cf. Sect. 9.2.4. However,
let us finally analyse the entropy balance, i.e. the second law of thermodynamics.
For a closed system it obeys

S2 − S1 = Sa,12 + Si,12 . (13.182)

Since the beer crate is assumed to be incompressible, the caloric equation of state
for entropy is as follows:
T2
S2 − S1 = mc ln = 0. (13.183)
T1

Hence, the generation of entropy is

Q 12
Si,12 = −Sa,12 = − < 0. (13.184)
T1

The entropy transferred by the heat can be easily calculated, since the temperature is
supposed to remain constant. Since thermal energy is supplied, Q 12 is positive. Due to

Si,12 < 0 (13.185)

the process is impossible, see Sect. 13.7.4.

Problem 13.5 Two identical vertical pipes with a diameter of D = 0.1 m are con-
nected by a thin tube and a valve, see Fig. 2.8. One pipe is filled with water up to a
height of H = 10 m. Water has a density of ρ = 1000 mkg3 , the other tube is initially
296 13 Meaning and Handling of Entropy

empty. Now the valve is opened. After some time, an equilibrium of the water quan-
tity is reached. Water is supposed to be an incompressible liquid with c = 4.18 kgkJK .
Is the process reversible? Calculate the generated entropy. The temperature in state
(1) is T1 = 293.15 K. Calculate the thermodynamic mean temperature for the change
of state (1) → (2). Note that no energy is exchanged with the environment.

Solution

The temperature rise T was previously calculated in Problem 12.3 by applying the
first law of thermodynamics for closed systems and by applying the caloric equation
of state. The temperature T2 of the new equilibrium state can therefore be calculated as

T2 = T + T1 = 293.1558 K. (13.186)

According to the entropy balance for closed systems it is

S2 − S1 = m (s2 − s1 ) = Si,12 + Si,12 . (13.187)

Since no energy is exchanged with the environment, it follows that

Sa,12 = 0. (13.188)

Thus, the generation of entropy yields

Si,12 = m (s2 − s1 ) . (13.189)

States ((1) and (2)) are unambiguously given. The caloric equation of state for the
specific entropy can be applied. For an incompressible fluid it is as follows

T2
Si,12 = m (s2 − s1 ) = mc ln . (13.190)
T1

Finally, the generation of entropy results in

J
Si,12 = 6.5706 > 0. (13.191)
K
Since the generation of entropy is positive, the process is possible but irreversible,
cf. Sect. 13.7.4. Once state (2) is reached, the system can therefore no longer move
back to state (1) without external influences from the environment.
Equation 13.173 can be applied to calculate the thermodynamic mean tempera-
ture, i.e.

12 mg H gH
Tm = = T
= = 293.1529 K. (13.192)
Si,12 4mc ln T1
2
4c ln TT21
13.7 Entropy Balancing 297

Fig. 13.24 Sketch to Problem 13.6

Problem 13.6 An ideal gas flows isobarically through an adiabatic horizontal chan-
nel. Its velocity decreases from c1 = 100 ms to c2 = 20 ms .
(a) Calculate the specific dissipated energy ψ12 .
(b) Does the temperature of the fluid increase, decrease or does it stay constant?
(c) Deduce a correlation to determine the thermodynamic mean temperature.
Solution

The problem is sketched in Fig. 13.24, that shows the first as well as the second law
of thermodynamics.
(a) The specific dissipated energy ψ12 can be derived from the partial energy equa-
tion, i.e.

2
1 2 
wt,12 = v d p + ψ12 + c2 − c12 + g (z 2 − z 1 ) . (13.193)
2
1

Since no technical work is transferred and the channel is horizontal the equation
simplifies to
2
1 2 
0 = v d p +ψ12 + c2 − c12 . (13.194)
2
1
  
y12 =0

The specific pressure work y12 is zero, since the flow is supposed to be isobaric,
i.e. p2 = p1 . Thus, the specific dissipation results in

1 2  J
ψ12 = − c − c12 = 4800 . (13.195)
2 2 Kg

(b) In order to gain information about the temperature, the first law of thermodynam-
ics is applied for an open system in steady state, ignoring the potential energy, i.e.
298 13 Meaning and Handling of Entropy

1 2 
q12 + wt,12 = (h 2 − h 1 ) + c2 − c12 . (13.196)
  2
=0 =0

With the caloric equation of state it follows

1 2 
0 = c p (T2 − T1 ) + c2 − c12 . (13.197)
2
Thus, the change of temperature is

1  2  ψ12
T = T2 − T1 = − c2 − c12 = > 0. (13.198)
2c p cp

The temperature rises.

Alternatively, the second law of thermodynamics can be applied, i.e.

Ṡa + Ṡi + ṁs1 = ṁs2 . (13.199)

Ṡa is zero, since the system is adiabatic. Ṡi must be positive, since dissipation
occurs. Thus, it is
ṁ (s2 − s1 ) = Ṡi > 0. (13.200)

Applying the caloric equation of state leads to

T2 p2 Ṡi
c p ln − R ln = > 0. (13.201)
T1 p ṁ
  1
=0

Equation
T2
c p ln >0 (13.202)
T1

can only be fulfilled in case T2 > T1 .


(c) It is known from Sect. 13.7.3 that the thermodynamic mean temperature obeys

˙ ψ12
Tm = = . (13.203)
Ṡi si,12

With Eq. 13.200 and the caloric equation of state it finally is

ψ12 ψ12 c22 − c12


Tm = = = − . (13.204)
s2 − s1 T
c p ln T21 2c p ln TT21
13.7 Entropy Balancing 299

Fig. 13.25 Left: Compression to the same volume, right: Expansion to the same pressure (both
adiabatic)

Problem 13.7 The adiabatic compression/expansion shown in Fig. 9.9 shall be plot-
ted in a T, s-diagram!
Solution

See Fig. 13.25.

Problem 13.8 An ideal gas G with a mass m G is trapped in a vertical cylinder C


(diameter d). The cylinder is closed by a freely movable, frictionless piston P. Piston
and cylinder shall be incompressible and made of the same steel. The masses of
cylinder and piston are given, see Fig. 13.26. In state (1) cylinder, piston and gas
have the same temperature T1 that is greater than ambient temperature Tenv . During
the change of state the entire systems cools slowly down until in state (2) ambient
temperature is reached. The change of state proceeds so slowly that piston, cylinder
and gas always have the same, homogeneous temperature. Changes of the potential
energy of the gas as well as the acceleration of the piston can be ignored. Calculate
for the change of state (1) → (2)
(a) the gas pressure,34
(b) the change of the potential energy of the system (cylinder, piston and gas),
(c) the entire released heat to the environment,
(d) the change of the entropy of the system (cylinder, piston and gas),
(e) the total generation of entropy (including the environment),
(f) the thermodynamic mean temperature!
Solution
(a) This case has already been treated in Example 7.6. Since the process is supposed
to be very slow, i.e. free of dissipation, the weight of the gas is neglected, the
piston moves without friction and the acceleration can be neglected, the change

34 Assume, that the weight of the gas is small compared to the mass of the piston.
300 13 Meaning and Handling of Entropy

Fig. 13.26 Sketch to Problem 13.8

of state is isobaric. Thus, an equilibrium of forces can be stated for the piston,
i.e.
p A = m P g + penv A (13.205)

with π 2
A= d = 0.0314 m2 . (13.206)
4
Hence, the pressure is
mPg
p= + penv = 1.6245 bar = const. (13.207)
A
(b) According to the system boundary, cf. Fig. 13.26, the entire change of potential
energy is

E pot,12 = E pot,12,P + E pot,12,G + E pot,12,C (13.208)


     
=0 =0

respectively
E pot,12 = m P g h (13.209)

with
h = h 2 − h 1 . (13.210)

Applying the thermal equation of state for states (1) and (2) leads to

pVG,1 = p Ah 1 = m G RT1 (13.211)


13.7 Entropy Balancing 301

Fig. 13.27 System boundaries for Problem 13.8

and
pVG,2 = p Ah 2 = m G RTenv . (13.212)

Hence, the change of height is

mG R
h = h 2 − h 1 = (Tenv − T1 ) = −1.9657 m. (13.213)
pA

Consequently, the total change of potential energy is

E pot,12 = m P g h = −3.8567 × 103 J. (13.214)

(c) Let us apply the first law of thermodynamics for system PGC as given by the
system boundary in Fig. 13.27a, i.e.

W12 + Q 12 = Utotal,12 + E pot,12 . (13.215)

Hence, the exchanged heat follows

Q 12 = Utotal,12 + E pot,12 − W12 . (13.216)

Applying the caloric equation of state for the internal energy results in

Utotal,12 = (m P + m C ) cS (Tenv − T1 ) + m G cv,G (Tenv − T1 ) . (13.217)


302 13 Meaning and Handling of Entropy

In order to determine the work W12 , that is exchanged with the environment,35
the partial energy equation is applied, i.e.

W12 = W12,V + env . (13.218)

With respect to the chosen system boundary, in this equation W12,V is at the
interface with the environment, so that the work can be formulated from the
perspective of the environment: It includes volume work at the environment and
dissipation in the environment env . Both can easily be calculated in order to find
W12 . Since the environment is supposed to be homogeneous,36 it follows that

env = 0. (13.219)

Thus, the partial energy equation simplifies to

2
W12 = W12,V = − penv dV = − penv (V2 − V1 ) = − penv A h > 0.
1
(13.220)
This makes sense because the piston moves downwards, i.e. the environment
expands so that work is supplied to the system. The transferred heat therefore
follows
 
Q 12 = (m P + m C ) cS + m G cv,G (Tenv − T1 ) + m P g h + penv A h.
(13.221)
In another notation this results in
 
Q 12 = (m P + m C ) cS + m G cv,G (Tenv − T1 ) +
mG R (13.222)
+ (m P g + penv A) (Tenv − T1 ) = −6.3193 × 107 J.
pA

(d) The entire change of entropy is composed of piston, cylinder and gas, see
Fig. 13.27a. In contrast to the gas, steel is supposed to be incompressible. Con-
sequently, the caloric equation of state leads to


Tenv Tenv penv
S2 − S1 = (m P + m C ) cS ln + m G c p,G ln − R ln (13.223)
T1 T1 p1

with
J
c p,G = cv,G + R = 5250 . (13.224)
kg K

35 When calculating the work, always make sure that it corresponds to the selected system boundary.
36 From now on, this is always the case.
13.7 Entropy Balancing 303

Hence, the entire change of entropy is

J
S2 − S1 = SPGC = −1.2768 × 105 . (13.225)
K
(e) Heat transfer requires a temperature potential, i.e. a discontinuity or devia-
tion from homogeneity, see Fig. 13.27a. This imperfection is the cause of the
generation of entropy in this particular problem, see Sect. 14.2 for a detailed
explanation. Thus, when asking for total entropy generation, this imperfection
must be included in the consideration, since the gas itself is reversibly com-
pressed by the slow change of state. To be precise: The initial temperature
spread between PGC and the environment is the driver for the change of state.
It causes heat transfer, which decreases over time as the temperature poten-
tial becomes smaller. Once the temperatures have equalised, the system can
not move back to its initial state. The change of state must therefore be irre-
versible.

Therefore, the system boundary has been slightly extended to the PGCext system
so that the temperature imperfection responsible for the generation of entropy
is also just part of the system see Fig. 13.27b. Anyhow, the additional mass
of the surrounding environment is so small, that it does not contribute to the
extensive change of entropy. Nevertheless, the goal is to catch the imperfection.
Consequently, the total change of entropy then results in

Stotal = SPGCext = SPGC + Senv


  
=0
(13.226)
Q 12
≡ Si,12,total + Sa,12,total = Si,12,total + .
Tenv

Thus, the generation of entropy results in

Q 12 J
Si,12,total = SPGC − = 8.7891 × 104 > 0 (13.227)
   Tenv K
see d)

Hence, Si,12,total > 0 proves that the process is irreversible.

Alternatively, the imperfection, i.e. the cause for generated entropy can be fur-
ther investigated, see Fig. 13.28. Therefore, the boundary, where the heat trans-
fer takes place, is balanced. However, the boundary should be so small that it is
almost massless,37 i.e. the entropy that goes in must also go out. The incoming
entropy from the environment38 is at constant ambient temperature, whereas the
outgoing entropy on the left is at a varying temperature. Furthermore, generation

37 Therefore, no storage of entropy is possible.


38 Actually, it is outgoing, since Q 12 < 0.
304 13 Meaning and Handling of Entropy

Fig. 13.28 Massless


boundary, Problem 13.8(e)

of entropy is assumed. Hence, the balance reads as39 :

2
Q 12 δQ
Si,12,total + = . (13.228)
Tenv T
1

The generated entropy results in

2
δQ Q 12
Si,12,total = − . (13.229)
T Tenv
1

Obviously, the differential heat δ Q is required. However, Eq. 13.222 leads to the
differential notation, i.e.

  mG R
δ Q = (m P + m C ) cS + m G cv,G dT + (m P g + penv A) dT. (13.230)
pA

Thus, for the generated entropy it follows

2
  1
Si,12,total = (m P + m C ) cS + m G cv,G dT +
T
1
(13.231)
2
mG R 1 Q 12
+ (m P g + penv A) dT − .
pA T Tenv
1

Finally, the generated entropy results in

  Tenv
Si,12,total = (m P + m C ) cS + m G cv,G ln +
T1
(13.232)
m G R Tenv Q 12 J
+ (m P g + penv A) ln − = 8.7891 × 104 > 0.
pA T1 Tenv K

39 Entropy in is equal to entropy out!


13.7 Entropy Balancing 305

(f) For calculating the thermodynamic mean temperature of the system PGC, the
following equation is applied
Q 12
Tm = . (13.233)
Sa,12,PGC

The heat Q 12 has already been calculated. In order to determine the entropy
carried by the heat it is
2
δQ
Sa,12,PGC = . (13.234)
T
1

For solving the integral, Eq. 13.222 is applied in differential notation, i.e.
  mG R
δ Q = (m P + m C ) cS + m G cv,G dT + (m P g + penv A) dT. (13.235)
pA
Thus, for the transferred entropy the integral in Eq. 13.234 is solved

2
  1
Sa,12,PGC = (m P + m C ) cS + m G cv,G dT +
T
1
(13.236)
2
mG R 1
+ (m P g + penv A) dT.
pA T
1

The solution results in


  Tenv
Sa,12,PGC = (m P + m C ) cS + m G cv,G ln +
T1
(13.237)
m G R Tenv J
+ (m P g + penv A) ln = −1.2768 × 105 .
pA T1 K

According to Eq. 13.233, the thermodynamic mean temperature finally obeys

Q 12
Tm = = 494.95 K. (13.238)
Sa,12,PGC

Problem 13.9 Exhaust gas, to be treated as an ideal gas with κ = 1.38 and R =
307.08 kgJK , expands adiabatically and reversibly in a turbine.40 Its initial state (1) is
given by ϑ1 = 585 ◦ C. State (2) is characterised by ϑ2 = 390 ◦ C and p2 = 0.981 bar.
Potential as well as kinetic energy can be ignored.

40The turbine is discussed in Chap. 17 in detail. However, it is a component, that releases work
while a fluid expands from p1 to p2 < p1 .
306 13 Meaning and Handling of Entropy

(a) Calculate the exhaust gas pressure at the inlet p1 .


(b) What is the thermodynamic mean temperature from (1) → (2)?
Solution
(a) The change of state runs adiabatically (δsa = 0) and reversibly (δsi = 0), i.e. it
is isentropic (ds = 0). Thus, it is

κ−1
κ
T1
p1 = p2 = 2.502 bar. (13.239)
T2

(b) According to the second law of thermodynamics it follows

δq δψ
ds = δsa + δsi = + . (13.240)
T T
Rearranged, it is
T ds = δq + δψ. (13.241)

Introducing a thermodynamic mean temperature Tm the integration leads to

2 2 2
Tm ds = δq + δψ. (13.242)
1 1 1

Finally, Tm obeys
q12 + ψ12
Tm = . (13.243)
s2 − s1

Since the change of state is adiabatic, i.e. q12 = 0, and reversible, i.e. ψ12 = 0,
the change of state is isentropic, i.e. s2 − s1 = 0. According to Eq. 13.243 the
thermodynamic mean temperature can not be calculated directly, i.e.

0
Tm → . (13.244)
0
To solve this problem, it is assumed that the change of state is adiabatic but
irreversible. In this case, the process is polytropic with a polytropic exponent n.
In the limiting case n → κ, the change of state is identical to the given, isentropic
conditions from (1) to (2), so that

q12 + ψ12 ψ12


Tm = lim = lim . (13.245)
n→κ s2 − s1 n→κ si,12

To calculate Tm the following equations need to be solved sequentially for the


polytropic change of state n → κ
13.7 Entropy Balancing 307


n−1
T2 p2 n
= (13.246)
T1 p1

The partial energy equation is

2

RT1 T2
wt,12 = v d p + ψ12 = n − 1 + ψ12 . (13.247)
n−1 T1
1

Applying the first law of thermodynamics for an adiabatic change of state, i.e.
q12 = 0, leads to
wt,12 = h 2 − h 1 = c p (T2 − T1 ) . (13.248)

Accordingly, the specific dissipation for n → κ results in




RT1 T2
ψ12 = c p (T2 − T1 ) − n −1 (13.249)
n−1 T1

The change of specific entropy for n → κ is




 n−κ T2
s2 − s1 = si,12 = cp − R ln (13.250)
n−1 T1

Solving Eqs. 13.246, 13.249 and 13.250 numerically for n → κ leads to

ψ12
Tm = lim = 756.47 K (13.251)
n→κ si,12

respectively
ϑm = 483.32 ◦ C. (13.252)

Problem 13.10 A vertical cylinder with an inner diameter of d = 0.25 m and a


height of H = 2 m contains a freely movable piston with a height of δK = 0.06 m.
The mass of the piston shall be m K = 23 kg. The space above the piston is filled with
an ideal gas A, with an initial pressure of pA,1 = 120 kPa. At the lower end of the
cylinder an ideal gas B is supplied: In front of the throttle the pressure of gas B and
its temperature are constant, i.e. pB = 700 kPa and ϑB = 15 ◦ C. By supplying gas
B the piston is lifted and gas A is compressed, see Fig. 13.29. The temperature of
the entire system is constant for the change of state, i.e. ϑ = ϑB = 15 ◦ C. Ambient
temperature is ϑenv = 15 ◦ C. Potential energies of gas A and B are to be ignored.
(a) Calculate the pressure pA,2 of gas A at the end of the compression, when there
is a mechanical equilibrium in front and behind the throttle.
308 13 Meaning and Handling of Entropy

Fig. 13.29 Sketch to


Problem 13.10

(b) At which height z 2 does the piston stop, see Fig. 13.29?
(c) Calculate the heat Q 12 , that is exchanged between the content of the cylinder
and the environment.
(d) How much entropy Si,12 is generated in total?
Solution
(a) In state (2) a mechanical equilibrium is reached, so that the balance of forces for
the piston leads to
pA,2 A + m K g = pB . (13.253)

The cross section A follows


π 2
A= d = 0.0491 m2 . (13.254)
4
Thus, the pressure pA,2 is

mKg
pA,2 = pB − = 6.954 × 105 Pa. (13.255)
A
(b) The change of state for gas A is isothermal, i.e.

pA,1 V1 = pA,2 V2 . (13.256)

Substitution of the volumes leads to

pA,1 A (H − δK ) = pA,2 A (H − z 2 − δK ) . (13.257)


13.7 Entropy Balancing 309

Solving for z 2 results in


 
pA,1
z 2 = (H − δK ) 1 − = 1.6052 m. (13.258)
pA,2

(c) The problem describes an open system in non-steady state. The first law of
thermodynamics then reads as

U2 − U1 + m K gz 2 = W12 + Q 12 + m B h B . (13.259)

This equation is structured as follows:


• The left hand side, i.e. U2 − U1 + m K gz 2 describes the change of state of
the system, see boundary in Fig. 13.29. The right hand side summarises what
causes the change of state.
• W12 = 0, since according to the system boundary there is no exchange of
work,41 such as a compression of the environment for instance.
• m B h B stands for the energy flux into the system due to the mass supply from
outside, see Sects. 9.2.8 and 11.3.1.
• Q 12 is the heat that is asked for.
Specifying of U2 − U1 results in

m c (T − T1 ) + m B u B,2 − 0 +m K gz 2 = Q 12 + m B (u B + pB vB ) .
 A v 2    
Gas A Gas B
(13.260)
Following a mass balance it is

m B,2 = m B . (13.261)

This leads to
 
Q 12 = m B u B,2 − u B +m K gz 2 − m B pB vB (13.262)
  
=0

respectively

Q 12 = m K gz 2 − m B pB vB = m K gz 2 − pB VB,2 . (13.263)

Finally, it is

Q 12 = m K gz 2 − pB Az 2 = −5.4795 × 104 J. (13.264)

41The evaluation of whether work is exchanged is from the perspective of the environment here.
The exchanged work would take place at the interface between the environment and the system.
Therefore, if no work is done at the environment, no work enters the system.
310 13 Meaning and Handling of Entropy

(d) The balance of entropy for the system follows

S2 − S1 = m A sA + m B sB,2 − 0 = Sa + Si + m B sB . (13.265)


     
=SA =SB

In other words, the change of entropy of the system S = SA + SB is caused
by generation of entropy Si at the throttle, by supply of heat Q 12 and by the flux
of entropy due to the supply of mass. Rearranging leads to
 
m A (s2 − s1 )A +m B sB,2 − sB = Sa + Si . (13.266)
     
pA,2
=−RA ln pA,1
=0

The change of state is isothermal, so that

2
δQ Q 12
Sa = = . (13.267)
T T
1

Thus, it finally is
Q 12 pA,2
Si = − − m A RA ln . (13.268)
T pA,1

Applying the thermal equation of state results in

Q 12 pA,1 V1 pA,2 J
Si = − − ln = 120.48 . (13.269)
T T pA,1 K

Problem 13.11 A piston with a mass of m K = 1200 kg and a cross section of A =


1.15 m2 is fixed in a cylinder, see Fig. 13.30. The space below the piston (volume
V1 = 0.23 m3 ) is filled with air ( p1 = penv = 1 bar). The piston can move freely and

Fig. 13.30 Sketch to


Problem 13.11
13.7 Entropy Balancing 311

is lifted by a distance of z = 0.5 m by filling the space below with air taken from a
pressurised vessel (volume VB = 0.1 m3 ). The air in this vessel has an initial pressure
of pB1 , so that a transient process takes place very slowly due to the connecting
throttle. The volume in the connecting pipes as well as the potential energy of the
gas can be ignored. Air is treated as an ideal gas with R = 0.287 kgkJK .
(a) Calculate the pressure p2 in the space below the piston, when the piston comes
to rest at z 2 = z 1 + z.
(b) What needs to be the initial pressure pB1 in the pressurised vessel, so that the
piston is lifted by z and comes to rest with the throttle fully open? The change
of state is supposed to be isothermal and the initial temperature of the entire
system is 20 ◦ C.
(c) How much heat needs to be exchanged?
(d) How much entropy is generated in this process?
Solution
(a) In thermodynamic equilibrium a balance of forces can be applied for the piston,
i.e.
penv A + m K g = p2 A. (13.270)

Thus, the pressure p2 follows


mKg
p2 = penv + = 1.1024 × 105 Pa. (13.271)
A
(b) The total mass remains constant so that the thermal equation of state can be
applied twice. In initial state it reads as

p1 V1 pB1 VB
+ = mI (13.272)
RT RT
and in the final state it follows
p2 (V2 + VB )
= m II . (13.273)
RT
It is
m I = m II (13.274)

and
V2 = V1 + z A. (13.275)

Solving for pB1 results in

p2 (V1 + z A + VB ) − p1 V1
pB1 = = 7.6764 × 105 Pa. (13.276)
VB
312 13 Meaning and Handling of Entropy

Fig. 13.31 Sketch to


Problem 13.11

(c) Applying the first law of thermodynamics for the system boundary shown in
Fig. 13.31 leads to

Q 12 − Wenv = 
U +E pot,K + E pot,G . (13.277)
  
=0 =0

According to Fig. 13.31 Wenv leaves the system, so that it has a negative sign. The
system releases work to the environment, which is compressed by the ascending
piston. Since the environment is homogeneous there is no dissipation. Thus, the
partial energy equation for the work reads as42

Wenv = Wenv,V + env = penv A z. (13.278)



=0

The heat then follows

Q 12 = penv A z + m K g z = 6.3386 × 104 J. (13.279)

(d) Both tanks are homogeneous for themselves, as the process is supposed to be
slow. However, at the interface of both systems, i.e. at the throttle, there is an
imperfection due to the pressure imperfection. Hence, entropy is generated at
this spot, see Sect. 14.4. However, the entropy balance for the entire system
reads as
S2 − S1 = Sa,12 + Si,12 (13.280)

respectively
Si,12 = S2 − S1 −Sa,12 . (13.281)
  
=S

42 From the perspective of the environment.


13.7 Entropy Balancing 313

Fig. 13.32 Sketch to


Problem 13.11

According to Fig. 13.32 the change of entropy for both tanks follows
   
TII p2 TII p2
S = m 1 c p ln − R ln + m B c p ln − R ln . (13.282)
TI p1 TI pB1

Since the change of state is isothermal, the following simplification results


p2 p2
S = −m 1 R ln − m B R ln . (13.283)
p1 pB1

With the masses


p1 V1
m1 = = 0.273 kg (13.284)
RT
and
pB1 VB
m2 = = 0.9124 kg (13.285)
RT
the entire change of entropy is

p2 p2 J
S = −m 1 R ln − m B R ln = 500.54 . (13.286)
p1 pB1 K

The entropy exchanged with the heat can be calculated easily since the change
of state is isothermal, i.e.
Q 12
Sa,12 = . (13.287)
T
The generated entropy finally is

Q 12 J
Si,12 = S − Sa,12 = S − = 284.31 > 0. (13.288)
T K
314 13 Meaning and Handling of Entropy

Fig. 13.33 Sketch to


Problem 13.11

Alternatively, the partial energy equation can be applied, see Fig. 13.33. For this
system boundary the work W12 is composed of work for lifting the piston and
for compressing the environment. Thus, the system releases the work

W12 = −m K g z − penv A z. (13.289)

Applying the partial energy equation43 results in

W12 = −m K g z − penv A z = WV,1 + WV,B + . (13.290)

The volumes follow


pB1 VB = m B RT = p2 VB2 (13.291)

respectively
pB1
VB2 = VB = 0.6964 m3 (13.292)
p2

and
m 1 RT
V2 = = 0.2086 m3 . (13.293)
p2

The isothermal volume works follow accordingly


p2
WV,1 = p1 V1 ln (13.294)
p1

and

43The work for lifting the piston and compressing the environment W12 is provided by the system.
The system changes its volume and causes dissipation inside. Mechanical work inside the system
does not play a role, since the piston is not part of the system and the potential energy of the gas
shall be ignored. This is covered by the partial energy equation.
13.7 Entropy Balancing 315

Fig. 13.34 Sketch to


Problem 13.12

p2
WV,B = pB1 VB ln . (13.295)
pB1

The dissipation then is

 = −m K g z − penv A z − WV,1 − WV,B = 8.3348 × 104 J. (13.296)

Since the process is isothermal the generated entropy is

 J
Si,12 = = 284.31 > 0. (13.297)
T K

Problem 13.12 The vertical temperature gradient dT dz


in an isentropic, dry atmo-
sphere shall be derived. Air is supposed to be ideal with R = 287 kgJK and κ = 1.4.
The gravitational acceleration is g = 9.81 sm2 and shall be constant.
Solution
• Alternative 1:

In a first step a differential44 balance of forces is derived, cf. Fig. 13.34. The balance
obeys
( p + d p) A + ρ Ag dz = p A. (13.298)

A simplification yields
d p = −ρg dz. (13.299)

The density can be substituted by the thermal equation of state, i.e.

dp g
=− dz (13.300)
p RT

44 Note that the atmosphere is compressible, compare with Problem 4.2.


316 13 Meaning and Handling of Entropy

Since the atmosphere shall be isentropic it is



κ−1
κ
p
T = T0 . (13.301)
p0

A combination of Eqs. 13.300 and 13.301 leads to


κ−1
dp gp0 κ
=− κ−1 dz. (13.302)
p RT0 p κ

Rearranging brings
g κ−1
p− κ d p = −
1
p κ dz. (13.303)
RT0 0

An integration of Eq. 13.304 reads as

1 g κ−1
p − κ +1 = −
1
p κ z+C (13.304)
− κ1 +1 RT0 0

respectively
κ κ−1 g κ−1
p κ =− p κ z + C. (13.305)
κ −1 RT0 0

Introducing the boundary conditions at the ground, i.e. z = 0 : p = p0 , leads to

κ κ−1
C= p0 κ . (13.306)
κ −1

Thus, it is

κ−1

p κ κ −1 gz κ
= − + . (13.307)
p0 κ RT0 κ −1

By combining with Eq. 13.301 yields

κ −1 g
T = T0 − z. (13.308)
κ R
Hence, the vertical temperature gradient reads as

dT κ −1 g K
=− = −9.766 (13.309)
dz κ R km
13.7 Entropy Balancing 317

• Alternative 2:

The first law of thermodynamics is applied for a differential piece of mass dm,
that is lifted from ground (1) to an upper position (2). Obviously such a system is
closed, so that
δw + δq = du + dea,pot . (13.310)

Since the system is adiabatic (δq = 0) and frictionless (δψ = 0), i.e. isentropic, it
is
δw = du + dea,pot . (13.311)

The partial energy equation follows

δw = δwV + δwmech + δψ = − p dv + dea,pot . (13.312)



=0

Hence, it is
R
− p dv = du = cv dT = dT. (13.313)
κ −1

According to the thermal equation of state it is

RT
v= (13.314)
p

respectively



∂v ∂v RT R
dv = dp + dT = − d p + dT. (13.315)
∂p ∂T p2 p

Inserting dv in Eq. 13.313 leads to




RT R R
− p − 2 d p + dT = dT. (13.316)
p p κ −1

A simplification results in
dp κ dT
= . (13.317)
p κ −1 T

A combination of Eqs. 13.300 and 13.317 reads as

g κ dT
− dz = (13.318)
RT κ −1 T

respectively
318 13 Meaning and Handling of Entropy

Fig. 13.35 Sketch to


Problem 13.13

g κ −1
dT = − dz. (13.319)
R κ
Thus, the temperature gradient finally follows

dT κ −1 g K
=− = −9.766 (13.320)
dz κ R km

Problem 13.13 A vertical cylinder whose piston is connected to a spring is filled


with air at a pressure of p1 . The temperature is T1 = 300 K, see Fig. 13.35. The
volume of the air within the cylinder is V1 = 0.035 m3 . In initial state the spring
has no residual tension. The air within the cylinder is heated as long as its volume
reaches V2 = 1.5 · V1 . The process is supposed to be reversible. The mass of the
piston is m P = 20 kg. The spring force follows Hooke’s law: FF = K F · x with a
spring constant of K F = 0.11 kN m
= const. The gas constant is R = 0.287 kgkJK and
the specific heat capacity is cv = 0.717 kgkJK . Ambient pressure is penv = 1 bar. The
diameter of the cylinder is d = 0.1 m. The potential energy of the contents of the
cylinder is negligible.
(a) What are the pressure p1 and the mass of the air m within the cylinder?
(b) Derive a correlation between the pressure within the cylinder and the distance x.
(c) What are the pressure p2 and the temperature T2 ?
(d) What is the work W12 that is transferred by the air within the cylinder?
(e) What is the heat Q 12 that is supplied to the content of the cylinder?
(f) Sketch the change of state in a p, v- as well as in a T, s-diagram.
(g) Calculate the thermodynamic mean temperature Tm,12 .
13.7 Entropy Balancing 319

Solution
(a) The pressure p1 follows from a balance of forces according to

mPg 4m P g
p1 = penv + = penv + = 1.2498 bar. (13.321)
A π d2
The pressure is calculated with the thermal equation of state, i.e.

p1 V1
m= = 0.0508 kg. (13.322)
RT1

(b) As soon as the piston starts moving, an additional spring force acts on the gas,
so that the balance of forces then becomes
mPg FF mPg KF · x
p (x) = penv + + = penv + + . (13.323)
A A A A
Thus, the pressure rises linearly with the distance the pistons moves.
(c) The initial height of the cylinder is

V1
h1 = = 4.4563 m. (13.324)
A
The final height in state (2) is accordingly

V2
h2 = = 1.5 · h 1 = 6.6845 m. (13.325)
A
Hence the piston moves by a distance of

x̃ = h 2 − h 1 = 2.2282 m. (13.326)

According to Eq. 13.323 the pressure in state (2) is

mPg K F · x̃
p (x̃) = penv + + = 1.5619 bar. (13.327)
A A
The temperature in state (2) then is

p2 V2
T2 = = 562.36 K. (13.328)
mR
(d) The partial energy equation reads as

2
W12 = W12,V + 12 + W12,mech = − p dV. (13.329)
   
=0 =0 1
320 13 Meaning and Handling of Entropy

The pressure p can be substituted according to Eq. 13.323, i.e.

2

mPg KF · x
W12 = − penv + + dV. (13.330)
A A
1

Rearranging brings

2 2 x2 =x̃
mPg KF · x
W12 = − penv dV − dV − A dx . (13.331)
A A 
1 1 x1 =0 =dV

Solving finally results in

mPg 1
W12 = − penv (V2 − V1 ) − (V2 − V1 ) − K F x̃ 2 = −2.4602 × 103 J.
A 2
(13.332)
(e) The first law of thermodynamics reads as

Q 12 + W12 = U + E a,12 . (13.333)


  
=0

Solving for the heat Q 12 yields

Q 12 = U − W12 = mcv (T2 − T1 ) − W12 = 1.2017 × 104 J. (13.334)

(f) The p, v- as well as the T, s-diagram, see Fig. 13.36, have been calculated
numerically by splitting the change of state into differential changes of state
pi−1 , vi−1 → pi , vi . Equation 13.323 has shown that the pressure follows the
distance x and thus the volume linearly. Therefore, the polytropic exponent can
be calculated for any differential sub-step, according to

ln ppi−1
ni = i
vi . (13.335)
ln vi−1

Figure 13.36a indicates that the polytropic exponent n for the change of state
(1) → (2) is not constant, since the spring force, i.e. the external impact, varies
by time. The specific entropy at any sub-step can be calculated according to the
caloric equation of state, i.e.

Ti pi
si = si−1 + c p ln − R ln . (13.336)
Ti−1 pi−1
13.7 Entropy Balancing 321

Fig. 13.36 Solution to Problem 13.13

The T, s-characteristic is shown in Fig. 13.36b. The initial specific entropy in


state (1) was arbitrarily set to zero.
(g) The entropy balance reads as

S2 − S1 = Sa + Si . (13.337)

Since the change of state is supposed to be reversible it simplifies to




T2 p2
Sa = S2 − S1 = m c p ln − R ln . (13.338)
T1 p1

Finally, the thermodynamic mean temperature is

Q 12 Q 12
Tm,12 = =   = 417.24 K. (13.339)
Sa m c p ln T1 − R ln pp21
T2

Problem 13.14 In an adiabatic, horizontal cylinder (d = 12 cm) with a friction-


less, freely movable, gas-tight piston air is filled in the compartments A and B, cf.
Fig. 13.37. Air is to be treated as an ideal gas with κ = 1.4 and a specific heat capac-
ity of c p = 1.004 kgkJK . Kinetic energy of the gas shall be ignored. The mass of the
piston shall be m P = 20 kg. At initial state the piston is given with the following
parameters:
322 13 Meaning and Handling of Entropy

Fig. 13.37 Sketch to


Problem 13.14

• Compartment A: VA,0 = 2 l, pA,0 = 8 bar, ϑA,0 = ϑ0 = 20 ◦ C


• Compartment B: VB,0 = 10 l, pB,0 = 1 bar, ϑB,0 = ϑ0 = 20 ◦ C
• Piston: ϑP,0 = ϑ0 = 20 ◦ C, cP = 0.47 kgkJK
In a first step, the system shall be free of any friction respectively dissipation.
Furthermore, compartments A and B shall be adiabatic. The fixing of the piston is
now released and a periodic oscillation starts.
(a) What is the maximum velocity the piston reaches?
(b) Plot velocity and position of the piston as a function of time!
Now, in a second step, there shall be friction between the moving piston and the
cylinder. Compartments A and B as well as the piston are still adiabatic, the gas shall
still be free of dissipation. The mass of the cylinder shall be ignored. Due to the
friction at the piston the oscillation finally stops.
(c) Plot the position of the piston x = f (t)!
(d) Plot the temperature of the piston as a function of time! What are the temperatures
ϑA , ϑB and ϑP , when the system comes to rest?
Finally, after a long period of time, the system shall slowly strive into a full thermo-
dynamic equilibrium, i.e. compartments A and B shall be diabatic.45 However, the
cylinder shall still be adiabatic to the outer environment.
(e) What is the temperature once equilibrium is reached?
Solution
(a) The piston is accelerated as long as the forces are imbalanced. The gas in com-
partment A expands, so that the piston is accelerated to the right, where the gas
in B is compressed and decelerates the piston. Thus, the acceleration stops, when
pressure pA equals pressure pB . At that state the maximum velocity is reached.
The changes of state in A and B are isentropic, since the gas shall be free of
dissipation and no heat is exchanged, i.e.

45 Thus, heat can pass compartments A and B at the interface to the piston.
13.7 Entropy Balancing 323

κ
κ
VA,0 VA,0
pA = pA,0 = pA,0 (13.340)
VA x π4 d 2

VA,0
respectively with L A,0 = π 2 = 0.1768 m
4d


κ
L A,0
pA = pA,0 . (13.341)
x

For compartment B it follows accordingly



κ
L B,0
pB = pB,0 (13.342)
L A,0 + L B,0 − x

VB,0
with L B,0 = π 2 = 0.8842 m. The acceleration comes to an end, when
4d


κ
κ
L A,0 L B,0
pA,0 = pB,0 (13.343)
x@cmax L A,0 + L B,0 − x@cmax

so that the velocity is maximised. The position x@cmax of the piston then is

L 2A,0 + L A,0 L B,0


x@cmax =   κ1 = 0.4976 m. (13.344)
L A,0 + L B,0 ppA,0
B,0

At this position the piston has been supplied with the work of the expanding
system A and has released the work for compression of system B. Systems A
and B only transfer volume work, since the piston is horizontal, kinetic energy
of the gas shall be ignored and there shall be no dissipation. Thus, it follows
   
WP = + |WA | − |WB | = + WV,A  − WV,B  (13.345)

respectively

κ−1 
pA,0 VA,0 L A,0
WP = − −1
κ −1 x@cmax

κ−1 
pB,0 VB,0 L B,0
− −1
κ −1 L A,0 + L B,0 − x@cmax
= 861.7328 J. (13.346)

Applying the partial energy equation for the incompressible, frictionless piston
yields
1  2 
WP = WV,P +Wmech,P + P = m P cmax − c02 . (13.347)
  2
=0 =0
324 13 Meaning and Handling of Entropy

The maximum velocity then is



2WP m
cmax = = 9.2830 . (13.348)
mP s

(b) Applying Newton’s second law for the piston brings

d2 x
Feff = m P a = m P . (13.349)
dt 2
The resulting force follows


κ 
π 2 L A,0 κ L B,0 π 2
Feff = ( pA − pB ) d = pA,0 − pB,0 d .
4 x L A,0 + L B,0 − x 4

=A
(13.350)
Hence, the differential equation for x = f (t) is

κ
κ 
L A,0 L B,0 d2 x
pA,0 − pB,0 A = mP . (13.351)
x L A,0 + L B,0 − x dt 2

The velocity of the piston then reads as

dx
c (t) = (13.352)
dt
respectively the acceleration
d2 x
a (t) = . (13.353)
dt 2
A numerical solution is plotted in Fig. 13.38.
(c) There is now supposed to be friction between the piston and the cylinder. A func-
tion is required to describe the frictional forces acting at the interface between
piston and cylinder. It is assumed, that such a force is proportional to the veloc-
ity of the moving piston, i.e. Ffric ∝ c. Hence, a proportional constant η > 0 is
introduced, so that
dx
Ffric = ηc = η . (13.354)
dt
Newton’s second law for the piston then reads as

d2 x
Feff − Ffric = m P a = m P (13.355)
dt 2
13.7 Entropy Balancing 325

Fig. 13.38 Sketch to


Problem 13.14

respectively46

κ
κ 
L A,0 L B,0 dx d2 x
pA,0 − pB,0 A−η = m P 2 . (13.356)
x L A,0 + L B,0 − x dt dt

By introducing the frictional force, cf. Eq. 13.354, it is ensured, that the friction
always acts against the direction of movement: In case the piston moves to
the right with respect to the chosen coordinate system, i.e. c > 0, the frictional
force contributes to a deceleration, i.e. −ηc < 0. In case the piston moves to
the left, i.e. c < 0 the frictional force also decelerates the piston, i.e. −ηc > 0.
Figure 13.39 shows a numerical solution of Eq. 13.356. Both cases (reversible,
i.e. η = 0, and irreversible, i.e. η > 0) are plotted.
(d) The first law of thermodynamics for the piston in differential notation yields

δW + δ Q = dU + dE a . (13.357)
 
=0 =m P c dc

Hence, it is
δW = dU + m P c dc. (13.358)

In case friction occurs, the partial energy equation for the incompressible piston
reads as

δW = δWV,P +δWmech,P + δP = m P c dc + δP . (13.359)


  
=0

46 However, the change of state for the gas in A and B still is isentropic, since compartments A
and B are adiabatic and there shall be no dissipation within the gas. The only friction occurs at the
interface piston/cylinder.
326 13 Meaning and Handling of Entropy

Fig. 13.39 Sketch to Problem 13.14

Combining Eqs. 13.358 and 13.359 brings

δP = dU = m P cP dTP . (13.360)

In other words, it is friction that is responsible for a rise in the temperature of


the piston. The friction can be calculated according to the frictional force, i.e.

dx
δP = Ffric dx = η dx. (13.361)
dt
Hence, it is
dx
η dx = m P cP dTP (13.362)
dt
respectively

2
dTP η dx
= . (13.363)
dt m P cP dt

The temperature rise as a function of time is shown in Fig. 13.40. Now, let us
calculate the temperatures of A, B and the piston, once the system comes to rest.
As explained at the beginning of this problem, the system comes to rest, when
the pressure in A equals the pressure in B. The change of state for the gas in A
and B is isentropic under the given premises, so that

κ
κ
L A,0 L B,0
pA,0 = pB,0 . (13.364)
xeq L A,0 + L B,0 − xeq

Hence, the position of the piston then is


13.7 Entropy Balancing 327

Fig. 13.40 Sketch to


Problem 13.14

L 2A,0 + L A,0 L B,0


xeq =   κ1 = 0.4976 m. (13.365)
L A,0 + L B,0 ppA,0
B,0

The pressure of A and B then is



κ
L A,0
pA,eq = pB,eq = pA,0 = 1.8794 bar. (13.366)
xeq

The temperatures of A and B follow



κ−1
pA,eq κ
TA,eq = TA,0 = 193.8014 K (13.367)
pA,0

respectively

κ−1
pB,eq κ
TB,eq = TB,0 = 351.0610 K. (13.368)
pB,0

Now, let us focus on the piston:

Alternative 1

It has been shown, that any temperature rise of the adiabatic piston is caused
by friction. So, let there be a reversible change of state first. Under these condi-
tions, the maximum velocity occurs when a pressure equilibrium is reached, cf.
Eq. 13.348. Now, friction is switched on. It has further been shown, that

x@cmax = xeq = 0.4976 m. (13.369)


328 13 Meaning and Handling of Entropy

By applying the partial energy equation from state (max) to state (eq) for the
piston it follows47

1  
W = 0 = WV,P +P + m P 0 − cmax
2
. (13.370)
 2
=0

In other words, the entire kinetic energy of the piston is consumed by the friction:

1
P = 2
m P cmax . (13.371)
2
Integration of (13.360) leads to

(eq)
1
δP = P = 2
m P cmax = m P cP TP . (13.372)
2
(max)

Finally, the temperature rise of the piston is


2
cmax
TP = = 0.0917 K. (13.373)
2cP

The temperature of the piston in state (eq)48 then is

TP,eq = TP,0 + TP = 293.2417 K. (13.374)

Alternative 2

Let us apply the first law of thermodynamics for the piston from state (0) to state
(eq), i.e.
1  
W0,eq + Q = m P cP TP + m P ceq
2
− c02 = m P cP TP (13.375)
    2   
=WP =0
=0

The work acting at the piston due to the compression/expansion of A and B has
already been calculated, cf. Eq. 13.346. Hence, the temperature rise follows

47 Effectively there is no work exchange from (max) to (eq), since the piston in both states does
have the same position, see Eq. 13.369, i.e. effectively no compression/expansion of the frictionless
gas in A and B.
48 Note that T =T
P,0 P,max , since maximum velocity is reached in reversible case, i.e. no friction can
cause a temperature rise from (0) to (max). However, once maximum velocity is reached, friction
is switched on.
13.7 Entropy Balancing 329

WP
TP = = 0.0917 K. (13.376)
m P cP
(e) Finally, the systems strives in a thermodynamic equilibrium due to heat transfer
between A, B and piston as long as a uniform temperature Tfinal is achieved. The
first law of thermodynamics for the entire system from state (eq) to state (final)
reads49 as

W + Q = 0 = Utotal + E a

=0
   
= m A cv Tfinal − TA,eq + m B cv Tfinal − TB,eq
 
+ m P cP Tfinal − TP,eq (13.377)

The masses of A and B are


pA,0 VA,0
mA = = 0.0190 kg (13.378)
RTA,0

and
pB,0 VB,0
mB = = 0.0119 kg. (13.379)
RTB,0

Hence, the final temperature then follows

m A cv TA,eq + m B cv TB,eq + m P cP TP,eq


Tfinal = = 293.15 K. (13.380)
m A cv + m B cv + m P cP

Moreover, a mechanical equilibrium occurs, i.e.

m A RTfinal m B RTfinal
pfinal = = . (13.381)
VA,final VA,0 + VB,0 − VA,final
     
= pA,final = pB,final

Solving for VA,final results in

m A VA,0 + m A VB,0
VA,final = = 0.0074 m3 . (13.382)
mA + mB

Thus, the equilibrium pressure is

m A RTfinal
pfinal = = 2.1667 bar. (13.383)
VA,final

49The system does not transfer any work respectively heat to the environment. Furthermore, the
system is in rest in states (eq) and (final).
330 13 Meaning and Handling of Entropy

13.8 Entropy—Conclusion

As an engineer, one wants to have an idea of the parameters one is calculating with.
With the introduction of entropy in the last sections, this has only been achieved
to a very limited extent: The variable has been derived mathematically, it has been
shown that it is a new, powerful state value, but a clear, concise idea, as we have of
temperature or pressure, is missing. Even if there is no “simple” visualisation, as a
reader or student one wishes that entropy could be described in one sentence and thus
gain an understanding. One starts looking for technical literature and then realises that
no one is able to say in a few words what entropy actually is. The easiest way is to get
involved with this variable and realise what this new state variable actually represents:
First of all, entropy can be exchanged across system boundaries. This happens in
two ways:
• convectively via an incoming or outgoing mass flow, so that an increase or decrease
of entropy inside the system is possible.
• via exchanged heat, which causes an increase or decrease in entropy inside the
system.
In addition, entropy can occur inside a system, e.g. during dissipative processes or
during balancing processes due to potential differences. Inside, entropy can only be
generated, but not destroyed. Entropy is therefore not a conservation variable, but
can be balanced like the other extensive state variables. In the previous sections it has
been shown that thermodynamic processes can be evaluated with this new state value.
Furthermore, the direction of thermodynamic processes can be described, which is
not possible with the mere law of conservation of energy, i.e. the first law of ther-
modynamics. According to the first law of thermodynamics, it would be possible,
for example, to place a hand on a hot plate so that the hand cools down and the
plate continues to heat up. Experience shows that this is never the case, but that heat
obviously follows the temperature gradient and there is a high probability that the
hand will be burnt. With the help of the new state value entropy, it is possible to
describe this process and the direction of heat transfer cf. Sect. 14.2.

Get accustomed to this new state value!


Order and disorder approach

Entropy can as well be regarded as order respectively disorder in a thermodynamic


system. This approach follows the statistical mechanics and is described in much
more details e.g. in [11].

Example 13.4 (Melting ice)


An ice cube is a well ordered crystal. Let the cube in state (1) consist of water at
0 ◦ C. However, it takes a specific heat supply of 333 kg
kJ
for melting the ice. In case
the change of state (1) to (2) is quasi-static and reversible the entropy in state (2)
reads as
mh m
S2 = S1 + . (13.384)
T
13.8 Entropy—Conclusion 331

Fig. 13.41 Sandcastle according to [2]

Thus, the entropy has increased, i.e.

S2 > S1 . (13.385)

Furthermore, the water molecules are in a disordered state compared with state
(1), since the crystal structure is dissolved.

Example 13.5 (Sandcastle) In his novel [2] describes the nature of entropy with
an example of a sandcastle at the beach, see Fig. 13.41. In state (1) each grain of
sand is in a well-defined, thus ordered, position and in thermodynamic equilibrium
with the environment. Let us focus on a grain of sand in state (1). Due to ambient
impacts and the gravitational field, the castle will soon be in a much more disordered
state with the grains of sand be distributed arbitrarily on the ground. In state (2) the
grain shall be in thermodynamic equilibrium with the environment. The first law of
thermodynamics for the isolated overall system50 reads as


n
W12 + Q 12 = 0 = (U2 − U1 )air + (U2 − U1 )sand + m i g (z 2 − z 1 )i . (13.386)
i=1
  
<0

The change in the potential energy of the sand grains is negative because the sand-
castle collapses. With the caloric equation of state it is


n
n
0 = m air cv,air (T2 − T1 ) + m i csand (T2 − T1 ) + m i g (z 2 − z 1 )i . (13.387)
i=1 i=1

Hence, the change of temperature is

50 The system is treated as isolated and the potential energy of the air is ignored.
332 13 Meaning and Handling of Entropy


n
m i g (z 2 − z 1 )i
i=1
T2 − T1 = − > 0. (13.388)

n
m air cv,air + m i csand
i=1

However, the partial energy equation follows51

W12 = 0 = W12,v +W12,mech + 12 (13.389)


  
=0

respectively

n
0= m i g (z 2 − z 1 )i + 12 . (13.390)
i=1

The dissipation then follows


n
12 = − m i g (z 2 − z 1 )i > 0. (13.391)
i=1

The second law of thermodynamics for the overall isolated system reads as

S2 − S1 = Sa,12 +Si,12 = (S2 − S1 )air + (S2 − S1 )sand > 0 (13.392)



=0

with
2
δ
Si,12 = > 0. (13.393)
T
1

The entropy from (1) to (2) hence has increased. The same results can be achieved
with the caloric equation of state,52 i.e.

T2
n
T2
S2 − S1 = (S2 − S1 )air + (S2 − S1 )sand = m air cv,air ln + m i csand ln > 0.
T1 i=1
T1
(13.394)
Note that T2 > T1 .

51 The system is supposed to have a constant volume.


52 Sand is treated as an incompressible solid.
13.8 Entropy—Conclusion 333

The number of possible disordered particle arrangements such as in state (2) is


large.53 The probability of obtaining a disordered state is therefore higher than finding
the system in an ordered state. Therefore, entropy can also be related to a measure of
the probability of a state occurring—entropy has increased from (1) to (2). Transient
processes strive towards disordered states while entropy increases.

53There are numerous possible states of how the particles are disordered when the sandcastle
collapses. But there is only one ordered state as in (1).
Chapter 14
Transient Processes

Irreversibility plays an important role in the evaluation of thermodynamic processes.


If friction occurs, a system cannot be operated reversibly. This has been discussed
in Sect. 7.3: A wire pendulum, for example, no longer reaches its initial position
after a certain number of oscillations. The amplitude decreases over time until the
pendulum finally stops in its rest position. This is due to dissipation, i.e. friction at the
suspension point, as well as interactions between the environment and the pendulum.
Consequently, kinetic energy is dissipated and released into the environment, from
where it cannot be supplied to the system again. Consequently, the state value entropy
increases in such a case due to dissipation. It has been shown that the rate of entropy
generation is a quantitative measure of the degree of irreversibility.

14.1 Mechanical Driven Process

So far, dissipation has been treated based on mechanical effects: A piston causes
friction with the cylinder walls, or turbulent eddies are generated in a fluid, for
example. All these mechanical effects have in common that they require motion.
This motion is triggered by a mechanical imbalance. A homogeneous fluid in rest
does not cause any irreversibilities. However, this chapter also introduces other effects
that can cause irreversibility.

Example 14.1 A wire pendulum, see Fig. 14.1, is further investigated on the basis
of the two laws of thermodynamics.

• At the beginning at t = 0 the system possesses potential energy, since the mass
m has been lifted in a gravitational field. In case the process is reversible, this
potential energy is converted into kinetic energy, that has its maximum in state (1).

© Springer Nature Switzerland AG 2022 335


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_14
336 14 Transient Processes

Fig. 14.1 Wire pendulum

First law of thermodynamics from (0) → (1)rev :

1  2 
W01rev + Q 01rev = mc (T1rev − T0 ) + mg (z 1rev − z 0 ) + mg c1rev − c02 .
      2
=U E
  
pot
E kin
(14.1)
Partial energy equation for W01rev :

W01rev = 0 = W01rev,mech + 01rev . (14.2)


  
=0

The work W01rev is zero, since no work crosses the system boundary. The mechan-
ical work can be substituted, i.e.

1  2 
W01rev = 0 = mg (z 1rev − z 0 ) + mg c1rev − c02 . (14.3)
2
Thus, the velocity in state (1) results in

c1rev = 2gz 0 (14.4)

A combination of Eqs. 14.3 and 14.1 leads to

Q 01rev = mc (T1rev − T0 ) . (14.5)

Let us assume, that pendulum and environment are in perfect thermal contact, i.e.

T1rev = T0 = Tenv (14.6)

Consequently, the pendulum would not release any thermal energy, i.e.

Q 01rev = mc (T1rev − T0 ) = 0. (14.7)


14.1 Mechanical Driven Process 337

Second law of thermodynamics:


The pendulum is treated as an incompressible solid, so that

T1rev
S1rev − S0 = mc ln = 0 = Si,01rev + Sa,01rev . (14.8)
T0

The entropy remains constant, because T1rev = T0 . Since the system is still adia-
batic, i.e. Sa,01rev = 0, it results in

Si,01rev = 0 (14.9)

This is thermodynamic proof that the process, if it runs frictionless, is reversible.

• Now let us investigate what happens, in case the process is irreversible, i.e. friction
occurs. Potential energy from state (1) is converted into kinetic energy and finally
into frictional energy, measurable by a temperature increase. But let us take step
by step:
First law of thermodynamics from (0) → (1)irrev :

1  2 
W01irrev + Q 01irrev = mc (T1irrev − T0 ) + mg (z 1irrev − z 0 ) + mg c1irrev − c02 .
      2
=U E
  
pot
E kin
(14.10)
Partial energy equation for W01irrev :

W01irrev = 0 = W01irrev,mech + 01irrev . (14.11)


  
>0

The work W01irrev is zero, since no work crosses the system boundary. The mechan-
ical work can be substituted, i.e.

1  2 
W01irrev = 0 = mg (z 1irrev − z 0 ) + mg c1irrev − c02 + 01irrev . (14.12)
2
Thus, the velocity in state (1) results in

01irrev
c1irrev = 2gz 0 − < c1rev (14.13)
m

A combination of Eqs. 14.12 and 14.10 leads to

Q 01irrev + 01irrev = mc (T1irrev − T0 ) . (14.14)

Let us assume, that pendulum and environment are in perfect thermal contact, i.e.
338 14 Transient Processes

T1irrev = Tenv . (14.15)

Consequently, the system would have to release heat, according to

Q 01irrev = −01irrev < 0 (14.16)

Second law of thermodynamics:


The pendulum is treated as an incompressible solid, so that

T1irrev
S1irrev − S0 = mc ln = 0 = Si,01irrev + Sa,01irrev . (14.17)
T0

It further is
Q 01irrev
Sa,01irrev = . (14.18)
Tenv

Note that to calculate the entropy transferred by the heat, the temperature within
the system at which the heat exceeds the system boundary is required. In this case,
see system boundary in Fig. 14.1, it is ambient temperature.1

Q 01irrev
Si,01rirev = −Sa,01irrev = − >0 (14.19)
Tenv

This is thermodynamic proof that when friction occurs, the process is irreversible.
Problem 14.1 Two adiabatic tanks A and B contain nitrogen, to be treated as an
ideal gas, in the following states:
– Tank A: V = 1 m3 , p = 22 bar, T1 = 300 K
– Tank B: V = 1 m3 , p = 2 bar, T1 = 300 K
The two tanks are connected to each other so that pressure balancing can take place
between the two tanks. During this process, no heat is exchanged with the environ-
ment.
(a) What is the entropy change of the entire system and the entropy generation
during this balancing process?
(b) What would be the entropy change if tank B were initially fully evacuated?

Solution

(a) The problem is sketched in Fig. 14.2. Although gas A and gas B are identical,
they are coloured differently to distinguish them for the following considerations.
The change of entropy results in

1 Due to the perfect thermal contact, the pendulum itself is always at ambient temperature.
14.1 Mechanical Driven Process 339

Fig. 14.2 Sketch to Problem 14.1

Stotal = S2 − S1 . (14.20)

The entropy of the two masses A and B can be superimposed, since entropy is
an extensive state value, i.e.

Stotal = (S2 − S1 )A + (S2 − S1 )B . (14.21)

Since the gas is supposed to be ideal, the caloric equation of state can be applied
as follows



T2 p2 T2 p2
Stotal = m A c p ln − R ln + m B c p ln − R ln (14.22)
TA pA TB pB

with
TA = TB = T1 = 300 K. (14.23)

In a next step temperature T2 and pressure p2 in the new equilibrium state (2)
need to be calculated. Let us start with the temperature T2 , that can be derived
from the first law of thermodynamics2 :

W12 + Q 12 = U2 − U1 = (U2 − U1 )A + (U2 − U1 )B . (14.24)


 
=0 =0

The work W12 is zero, since there is no work being transferred across the system
boundary. Once again, the caloric equation of state can be applied, i.e.

2 Note that the internal energy is an extensive state value and can therefore also be superimposed.
Furthermore, there is no change in the potential energies because the gas occupies the entire volume
of the tanks. Therefore, the vertical position of the centre of gravity is fixed at half the height of the
tanks.
340 14 Transient Processes

0 = m A cv (T2 − T1 ) + m B cv (T2 − T1 ) . (14.25)

This equation implies, that the change of state needs to be isothermal,3 i.e.

T2 = T1 . (14.26)

In order to calculate the pressure p2 the thermal equation can be applied:


• Single systems A and B in state (1)

pA V
pA VA = m A RT1 ⇒ m A = (14.27)
RT1

pB V
pB VB = m B RT1 ⇒ m B = (14.28)
RT1

• Entire system in state (2)

p2 2V = (m A + m B ) RT2 (14.29)

With Eqs. 14.27 and 14.28 it results in

(m A + m B ) RT2 pA + pB
p2 = = = 12 bar. (14.30)
2V 2
Since we now know T2 and p2 , the entropy change can be calculated according
to Eq. 14.22, i.e.
p2 p2
Stotal = −m A R ln − m B R ln . (14.31)
pA pB

With Eqs. 14.27 and 14.28 it results in

pA V p2 pB V p2 J
Stotal = − ln − ln = 3250 . (14.32)
T1 pA T1 pB K

The following step clarifies how this entropy change is triggered. In general, the
entropy change is caused by entropy transported with the heat and by internally
generated entropy, i.e.
Stotal = Sa,12 +Si,12 . (14.33)

=0

Sa,12 needs to be zero since the system is adiabatic, i.e.

3 However, in the similar Problem 11.2 respectively Problem 12.3 there is a rise of temperature due
to the change of potential energy.
14.1 Mechanical Driven Process 341

J
Si,12 = Stotal = 3250 > 0. (14.34)
K
Hence, the change of state is possible, but irreversible, cf. Sect. 13.7.4. Once,
the thermodynamic equilibrium is reached, the system can not drive itself back
into initial state without any external impact from outside. However, applying
the partial energy equation would lead to the same result4 :

W12 = 0 = W12,V,A + W12,V,B + W12,mech +12 . (14.35)


  
=0

Thus, it is
12 = −W12,V,A − W12,V,B . (14.36)

Since the change of state is isothermal, the volume work can be calculated easily
and follows p2 p2
12 = − pA V ln − pB V ln . (14.37)
pA pB

Under these isothermal conditions it is


12 pA V p2 pB V p2 J
Si,12 = =− ln − ln = 3250 . (14.38)
T1 T1 pA T1 pB K

(b) The same calculation as in (a) can be performed when system B is fully evacuated,
i.e. m B = 0. The total change of entropy follows accordingly, i.e.


T2 p2
Stotal = m A c p ln − R ln . (14.39)
TA pA

With the same considerations as in (a) it follows, that

T2 = T1 (14.40)

and
1
p2 = p1 = 11 bar. (14.41)
2
This leads to
pA V p2 J
Stotal = − R ln = Sa,12 +Si,12 = 5080 > 0. (14.42)
T1 pA  K
=0

Obviously, the pressure difference between the two tanks A and B in case (b) is
greater than in case (a). Consequently, the balancing process is more intense in

4 W12 = 0, since the system does not exchange any energy with the environment.
342 14 Transient Processes

Fig. 14.3 Wall—Heat conduction

case (b) than in case (a). In conclusion, the entropy generated is a measure of
the intensity of the balancing process.

14.2 Thermal Driven Process

As already mentioned, the phenomena that lead to generation of entropy have so


far been based on mechanical effects. The driving force for this mechanism is the
movement of a fluid or a component, e.g. a piston. However, in systems driven,
for example, by thermal imbalances, there is a different mechanism for entropy
generation.
Conductive wall
To understand these thermal effects, let us consider a simple example shown in
Fig. 14.3. A solid wall separates the interior of a house from the environment. Inside
the house, the temperature shall be higher than the ambient temperature. Let us
assume that the surface temperature of the wall is T1 on the inside and T2 on the
outside. From a thermodynamic point of view, this is a thermal imbalance and the
system has the tendency to reach a new state of equilibrium. Hence, a heat flux is
triggered from the inside to the outside.5 Obviously, the wall is mechanically at per-
fect rest, so there is no dissipation due to motion as known from systems treated
before.

According to the first law of thermodynamics, in steady state the incoming energy
fluxes are equal to the outgoing energy fluxes, i.e.

Q̇ in = Q̇ out = Q̇. (14.43)

5Everyday experience shows, that a house needs a heating system in order to compensate the heat
being released to environment.
14.2 Thermal Driven Process 343

Otherwise, the temperature of the wall would vary over time. The heat is accompanied
by an entropy flux, i.e.

Ṡa = . (14.44)
T
T is the temperature at the system boundary within the system, where the heat passes.
Due to T1 > T2 less entropy enters than entropy leaves:

Q̇ Q̇
< . (14.45)
T1 T2

In a steady state, however, the incoming entropy must be equal to the outgoing
entropy. If this is not the case, the entropy in the system would vary in time. Since
entropy is a state value, the state itself would vary temporally, which is contradictory
to the steady state. In order to solve this dilemma, entropy must be generated within
the system, i.e. Ṡi > 0:
Q̇ Q̇
+ Ṡi = . (14.46)
T1 T2

Hence, the generated entropy follows

Q̇ Q̇
Ṡi = − . (14.47)
T2 T1

In a different mathematical notation it results in

T1 − T2
Ṡi = Q̇ ≥0 (14.48)
T1 T2

due to
T1 ≥ T2 . (14.49)

Theorem 14.1 Entropy is generated at the imperfection, i.e. the wall impedes the
balancing process between inside and outside. In contrast, in homogeneous systems,
i.e. in systems without an imperfection, no entropy is generated.
It is known, that Ṡi can never be negative, i.e. entropy can not be destroyed, see
Sect. 13.7.4. This leads to the conclusion, that the direction of heat transfer, as
assumed in Fig. 14.3, can only occur, if T1 ≥ T2 :
Theorem 14.2 Heat always follows the temperature gradient from warm to cold.
According to Eq. 14.48 the generation of entropy rises with increasing temperature
difference. Temperature differences lead to a balancing process. Once again,6 the
entropy generated is a measure of the intensity of a process.

6 Comparable to the pressure balancing in Problem 14.1.


344 14 Transient Processes

Fig. 14.4 Heat transfer


between two homogeneous
systems

Heat transfer between two systems


This example expands the view of balancing processes. A distinction is made between
homogeneous systems and imperfections, i.e. inhomogeneous systems, cf. Fig. 14.4.
Consider the heat transfer between two systems A and B, both of which are
adiabatic to the environment. The problem is guided by the following premises:
• Both systems are separated from each other by a wall, i.e. there is no mass flux.
They have different, but uniform temperatures with

TB > TA . (14.50)

This temperature potential triggers a heat flux between systems B and A, as the
overall system strives for thermal equilibrium.
• The systems are supposed to be large, i.e. the temperatures remain constant in
time, although thermal energy leaves systems A and enters system B.
• Thus, each system is homogeneous. There are no imperfections within the systems.
Consequently, there is no generation of entropy within system A respectively
system B.
However, the overall process is irreversible, as heat is transferred between two sub-
systems with different temperatures. This is explained in this section.
The change of entropy in system B follows7 :

δQ
dSB = − + δSi,B with δSi,B = 0 (homogeneous phase). (14.51)
TB

The change of entropy in system A follows:

δQ
dSA = + δSi,A with δSi,A = 0 (homogeneous phase). (14.52)
TA

Hence, the overall change of entropy, as extensive state value, is composed of systems
A and B:

7 See Theorem 14.1.


14.2 Thermal Driven Process 345

dStotal = dSA + dSB = δSa,total + δSi,total = δSi,total . (14.53)


  
=0,adiabatic

Obviously, the overall change of state is due to internal heat transfer8 and due to
generation of entropy. There are no influences respectively impacts from outside.
A combination of Eqs. 14.51–14.53 results in

δQ δQ
δSi,total = dSA + dSB = − . (14.54)
TA TB

In a different notation it reads as

TB − TA
δSi,total = δ Q >0. (14.55)
TA TB

Thus, the conclusion is


• The requirement of the second law of thermodynamics (δSi,total > 0) is fulfilled if
TB > TA . From Theorem 14.2 it follows that heat is always transferred from hot
to cold.
• At the interface between systems A and B there is irreversible entropy generation.
There is a discontinuity at this interface and the temperature profile is inhomoge-
neous.
• The energy of the overall system remains constant, since no external influences
cross the system boundary. However, entropy inside the fully-closed system rises
due to the thermal balancing process.
• Entropy generation increases with increasing temperature difference, i.e. with ther-
mal potential.
• There is a limiting case of reversible heat transfer when the temperatures of system
A and B are asymptotically equal. In order to still transfer heat, the effective surface
area must be very large under these circumstances. This can be shown with the
following equation9 :
Q̇ = k A T. (14.56)

Example 14.2 This example refers to the temperature measurement according to the
zeroth law of thermodynamics, see Fig. 4.5. The corresponding balancing process
is now described thermodynamically. Figure 14.5 shows the first and second law of
thermodynamics as well as a temperature profile. Let us assume that the temperature
of system C is constant,10 even if the thermometer, i.e. system A, is in contact with
C. However, thermal balancing takes place and the temperature of A changes with
time. The temporal heat flux is known from Heat and Mass Transfer, i.e.

8 In this case δSa,total is zero, since the entire system shall be adiabatic to the environment.
9 This equation is derived in the lecture Heat and Mass Transfer.
10 Due to the size of C.
346 14 Transient Processes

Fig. 14.5 Thermal balancing process

Q̇ (t) = k A [TC − TA (t)] . (14.57)

Under the premise that the temperature of thermometer A is uniform, the first law of
thermodynamics obeys
dUA TA
= m A cA = + Q̇ (t) . (14.58)
dt dt
Thus, it is
TA
m A cA = k A [TC − TA (t)] . (14.59)
dt
This differential equation can be solved by separating the variables and integration,
i.e.
dTA kA
=− dt. (14.60)
TA − TC m A cA

By applying the initial condition, cf. Fig. 4.5, it follows

TA (t = 0) = TB . (14.61)

Temperature TA (t) results in

− m k Ac t
TA (t) = TC − (TC − TB ) e A A (14.62)

The heat flux follows from Eq. 14.57, i.e.

− m k Ac t
Q̇ (t) = k A (TC − TB ) e A A (14.63)
14.2 Thermal Driven Process 347

These functions are plotted in Fig. 14.5. Obviously, the temperatures equalise after a
long time. At the beginning, the temperature potential TC − TA is maximum. Conse-
quently, the heat flux is also maximum at the beginning and decreases as time t pro-
gresses. Let us now examine the consequences for entropy. A system boundary is cho-
sen exactly at the intersection of systems A and B. Such a system is massless, i.e. no
entropy accumulation takes place.11 The incoming entropy flux due to the heat flux is

Q̇ (t)
Ṡa,in (t) = . (14.64)
TC

Accordingly, the outgoing entropy flux results in

Q̇ (t)
Ṡa,out (t) = . (14.65)
TA (t)

As long as TC > TA the incoming entropy flux is smaller than the outgoing flux.
Since no entropy accumulation is possible due to the massless system, i.e. the inter-
face between A and C, the following applies

Ṡin = Ṡout . (14.66)

Thus, in order to balance incoming and outgoing entropy flux, a flux of generation
of entropy Ṡi is required

Ṡa,in (t) + Ṡi (t) = Ṡa,out (t) . (14.67)

For the flux of generation of entropy it is

Q̇ (t) Q̇ (t)
Ṡi (t) = − . (14.68)
TA (t) TC

This leads to:


 
− m k Ac t 1 1
Ṡi (t) = k A (TC − TB ) e A A − (14.69)
− m k Ac t TC
TC − (TC − TB ) e A A

As soon as the temperatures A and C are equalised, the entropy generation disap-
pears. The larger the temperature spread between A and C is, the more entropy is
generated. The generation of entropy occurs at the imperfection at the A/C interface,
which is indicated by a temperature difference. When this imperfection disappears,
the generation of entropy stops, see Fig. 14.5. This is at t → ∞.

11Furthermore, the heat flux released by C is fully supplied to A at such a massless interface, since
no energy accumulation takes place.
348 14 Transient Processes

14.3 Chemical Driven Process

Chemical driven processes are thermodynamically investigated in Parts II and III.

14.4 Conclusions

In Chap. 13 the generation of entropy has been investigated. In any mechanical sys-
tem, internal friction, i.e. dissipation, may occur due to the movement of mechanical
parts, e.g. a piston moving in a cylinder, or due to the motion of a fluid.12 In such
processes energy is dissipated, indicated by ψ12 > 0. This leads to generation of
entropy, i.e. si,12 > 0. In Chap. 16 exergy is introduced: it is shown, that generation
of entropy reduces the working capability of a system. In thermodynamics, however,
there is a limiting case in which systems are free of any dissipation. This is a theoret-
ical benchmark for systems that operate reversibly. Such systems achieve the highest
efficiency, cf. Chap. 15. Nevertheless, these ideal systems do not exist in reality since
dissipation can not be prevented.

In addition to the mechanically caused generation of entropy, there is another


mechanism of non-homogeneous systems characterised by imperfections: In these
systems, entropy is generated at imperfections, see Sect. 14.2.
Theorem 14.3 Any disturbance of the thermodynamic equilibrium13 triggers a bal-
ancing process to achieve a new state of equilibrium, i.e. mechanical, thermal and
chemical equilibrium. This balancing process is irreversible and characterised by
generation of entropy.
Examples of transient balancing processes due to mechanically imbalanced sys-
tems are:
• pressure, see Problem 14.1
• or height, see Problem 13.5
Thermal imbalanced system are characterised by a heterogeneous temperature distri-
bution. Chemical imperfections are due to a concentration imbalance. In Sect. 3.2 it
has been shown, that thermodynamic systems always strive for mechanical, thermal
and chemical equilibrium.
Theorem 14.4 The greater the imperfection, the more entropy is generated. Homo-
geneous systems, that do not have imperfections, i.e. gradients, do not produce
entropy.

12 This can be in open as well as in closed systems.


13 Pressure, temperature and chemical potential.
14.4 Conclusions 349

Fig. 14.6 Sketch to Problem 14.2

Once the mechanical, thermal or chemical potential has been equalised by a balancing
process, the system can no longer move back to its initial state without external
impact.14

Problem 14.2 Through the wall of a house with an internal temperature of ϑ =


22 ◦ C, a heat flux of Q̇ = 20 kW is released to the environment with a temperature
of ϑenv = −5 ◦ C. What is the entropy flux Ṡi , that is generated in this process?

Solution

The problem is sketched in Fig. 14.6. Obviously, in steady state, the heat flux entering
the system is equal to the heat flux leaving the system. In addition, all state values
in the system must be constant in time, so that the entropy in the system is also
constant, i.e.
dS
= 0. (14.70)
dt
Thus, the entropy flux into the system must be balanced by the entropy flux out.
According to Fig. 14.6 it is:
Q̇ Q̇
+ Ṡi = . (14.71)
T Tenv

Note that generation inside the system Ṡi is a source term. Finally, the generation of
entropy results in

1 1 W
Ṡi = Q̇ − = 6.82 > 0. (14.72)
Tenv T K

14 A car does not re-climb a hill without external impact, once it has reached its rest position. A cup
of tea does not heat up, once it has reached ambient temperature. Components in a gaseous mixture
do not segregate once they are perfectly mixed.
350 14 Transient Processes

The process is therefore irreversible, i.e. once the heat flux has left the building, it
does not return.

Problem 14.3 An ideal gas (R = 0.280 kgkJK ) with a mass of m G = 1 kg and a tem-
perature of ϑG = 20 ◦ C shall be reversibly and isothermally compressed from an
initial pressure p1 = 100 kPa to a final pressure p2 = 600 kPa. The cylinder is in
thermal contact with a heat reservoir of constant temperature.15
Calculate the change of entropy for the gas in the cylinder and for the heat reservoir, if
(a) the temperature of the reservoir is ϑR = ϑG = 20 ◦ C (Case 1)?
(b) the temperature of the reservoir is ϑR = 15 ◦ C (Case 2)?
(c) What is the generated entropy Si,12 in both cases?

Solution

(1) Let us start with the change of entropy for the gas, as illustrated by the system
boundary “G” in Fig. 14.7. Independently from case 1 or case 2 there are two
alternatives to determine SG :

• Alternative 1—Caloric equation of state

SG = (S2 − S1 )G = m G (s2 − s1 )G . (14.73)

Applying the caloric equation of state for an isothermal change of state results
in

Fig. 14.7 Sketch to Problem 14.3


15 The heat reservoir is supposed to be huge.
14.4 Conclusions 351


TG,2 p2 p2 kJ
SG = m G c p ln − R ln = −m G R ln = −0.50169 .
TG,1 p1 p1 K
(14.74)
• Alternative 2—Entropy balance

The entropy change of the gas is composed of the entropy that passes the
system boundary due to heat transfer and the entropy that is generated in the
system:
SG = Sa,12 + Si,12 . (14.75)

Since the gas is reversibly compressed, no entropy is generated in the gas as a


result of the compression., i.e. Si,12 = 0. Consequently, the change of entropy
obeys
2
δQ Q 12
SG = Sa,12 = = . (14.76)
T TG
1

Note that in this equation T is the temperature inside the system, where the
heat passes the boundary. In this case, since the system “G” is regarded, it is the
isothermal temperature TG . In a next step, the heat Q 12 needs to be calculated.
Thus, the first law of thermodynamics for system “G” is applied, i.e.

Q 12 + W12 = (U2 − U1 )G = m G cv (T2 − T1 )G = 0. (14.77)

Hence, it is
Q 12 = −W12 . (14.78)

The partial energy equation for W12 in reversible operation reads as16
p2
W12 = WV,12 + 12 = WV,12 = m G RTG ln . (14.79)
 p1
=0

Consequently, the heat follows


p2
Q 12 = −m G RTG ln = −147.071 kJ. (14.80)
p1

Finally, the change of entropy obeys

Q 12 kJ
SG = = −0.50169 . (14.81)
TG K

16 See Fig. 13.13.


352 14 Transient Processes

Both alternatives lead to the same result. The result is independent from the heat
reservoir, since only system “G” has been regarded.

(2) Now, let us focus on the heat reservoir “R”. A caloric approach as for the gas
(see alternative 1), i.e.

SR = (S2 − S1 )R = m R (s2 − s1 )R , (14.82)


   
→∞ =0

is not expedient, since the mass of the heat reservoir is huge and its state constant.
However, a balance of entropy leads to

SR = Sa,12 + Si,12 . (14.83)



=0

Since the heat reservoir is supposed to be huge and homogeneous, there is no


generation of entropy in the reservoir, i.e. Si,12 = 0. Since Q 12 leaves system
“R” at TR , it is
Q 12
SR = − . (14.84)
TR

Thus, according to case 1 and 2 it finally is



0.50169 kJ @ ϑR = 20 ◦ C
SR = K (14.85)
0.510398 K @ ϑR = 15 ◦ C.
kJ

(3) Entire system (extended system boundary)

According to the extended system boundary shown in Fig. 14.8 the total change
of state results in

Stotal = SG + SR + SCylinder = Si,12 + Sa,12 . (14.86)


   
=0 =0

The cylinder is ignored as it is considered an incompressible, isothermal body.17


The entire system is huge, so that it is adiabatic,18 i.e. Sa,12 = 0. Thus, the equa-
tion simplifies to
Stotal = Si,12 = SG + SR . (14.87)

According to cases 1 and 2 it finally is

17Hence, its state is constant.


18Due to its homogeneity, the outer edge of the environment has no temperature gradients, so that
no heat is transferred. Furthermore, Q 12 is within the system, i.e. it does not occur at the system
boundary.
14.4 Conclusions 353

Fig. 14.8 Sketch to Problem 14.3


0 kJ @ ϑR = 20 ◦ C
Si,12 = K (14.88)
0.0087 kJ
K
@ ϑR = 15 ◦ C.

In summary, in case 1, no entropy is generated because the gas is reversibly


compressed, i.e. no dissipation occurs, and the heat transfer takes place in the
limiting case of T = 0. This is a theoretical benchmark, as a huge area would
be needed for heat transfer, cf. Eq. 14.56. However, the entire system is homo-
geneous, as indicated by the temperature profile. The overall change of state is
reversible, the heat transfer could theoretically run in the opposite direction.19

In contrast, case 2 is irreversible, though there is no dissipation within the gas as


well. The reason for the irreversibility is the heat transfer, which takes place with
T = 5 K. The overall system is therefore not homogeneous, but characterised
by an imperfection, see the overall temperature profile for case 2 in Fig. 14.8. This
temperature step makes the entire change of state irreversible, as it is impossible
for the given system to initiate a heat transfer from cold to warm.

19 Due to the large heat transfer area, that allows heat transfer with T → 0.
Chapter 15
Second Law of Thermodynamics

Although the entropy equilibrium has been referred to as the Second Law of Ther-
modynamics in the previous chapters, its classical formulation goes hand in hand
with thermodynamic cycles. Therefore, the focus is now on these cycles. However,
there are two different types of thermodynamic cycles: Clockwise cycles, on the
one hand, convert heat into mechanical energy and are denoted as thermal engines.
Counterclockwise cycles, on the other hand, are refrigerators and heat pumps that
lift thermal energy from a lower to a higher temperature level. Initially, however,
the cycles are represented in a so-called black-box notation, i.e. the fluxes at their
boundary are accounted for without focusing on the detailed physical processes that
take place inside the machine. Finally, a detailed description of which components
these machines consist of is given in Part II.

15.1 Formulation According to Planck—Clockwise Cycle


Processes

According to Planck, a permanently or cyclically running machine1 that takes heat


from a reservoir and converts it exclusively into work is impossible, cf. Fig. 15.1. A
fictitious machine that would be able to do this violates the second law of thermo-
dynamics and is called Perpetuum Mobile of the second kind.2
According to the first law of thermodynamics, however, there is no contradiction:
for closed systems in steady state, see Fig. 15.1a, the energy supplied is balanced by
the energy leaving the machine, i.e.

1 Such a machine would be operated in steady state.


2 A fictitious machine that violates the first law of thermodynamics, i.e. releases more energy than
is supplied, is called a Perpetuum Mobile of the first kind.
© Springer Nature Switzerland AG 2022 355
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_15
356 15 Second Law of Thermodynamics

Fig. 15.1 Contradiction to


the second law according to
Planck

Q = W. (15.1)

Considering the law of conservation of energy, such a machine would therefore exist.
Thermal energy would be converted exclusively into mechanical work. In steady
state, however, each state value, e.g. pressure p, temperature T and internal energy
U , must be constant over time, although it may vary locally. Since entropy S is also a
state value, it must also be constant over time.3 Consequently, the incoming entropy
must be equal to the outgoing entropy, see Fig. 15.1b:

Q
+Si = 0. (15.2)
T

>0

This results in
Q
Si = − < 0. (15.3)
T
Entropy is carried into the machine by the heat supplied. Entropy can be generated
within the machine due to losses or imperfections, but entropy can not be destroyed.
However, no thermal energy is released, so no corresponding entropy is released.
According to Sect. 13.7.4, the process is impossible. The amount of entropy in the
system cannot be constant over time because entropy enters with the heat supplied
and entropy may be generated in the system while no entropy leaves the system.
Theorem 15.1 A machine that is operated in a steady state can never completely
convert thermal energy into mechanical energy.

Example 15.1 The following example shows how a steady-state thermal engine,
which is to convert thermal energy into mechanical energy, can be realised. It has

3 According to Fig. 15.1a, the state value temperature would remain constant, since the incoming
energy corresponds to the outgoing energy. However, the entropy as a state variable would increase
due to the supplied heat. Since, for example, for ideal gases the entropy is a function of temperature
and specific volume, the specific volume would have to change over time while the temperature
remains constant—the state would therefore not be stationary. In order to keep entropy constant,
entropy would have to be destroyed in the machine. However, this is not possible.
15.1 Formulation According to Planck—Clockwise Cycle Processes 357

Fig. 15.2 Thermal engine


(closed system) without
cooling

already been presented in Sect. 11.2.1. Thus Fig. 15.2 actually means opening the
black box, as shown in Fig. 15.1 to get a view into the system. A gas can be compressed
and it can expand. In order to release more work during expansion than work supplied
during compression, the expansion must take place at a greater pressure level. If this is
not realised, the changes of state run from (1)→(2)→(1) and expansion respectively
compression work have the same amount.
Therefore, after reaching the minimum volume, the pressure must be increased
by supplying thermal energy, see state (3). Now more work is released during the
expansion of the gas until state (4) is reached than is supplied for compression. Since
the machine is to run permanently, i.e. in steady state, compression is performed
again and state (3) is reached again.
Now, when heating up again, a higher pressure and temperature are reached than
in (3). If this cycle is repeated continuously, the temperature and pressure at bottom
dead centre rise continuously without releasing any effective work. All the thermal
energy supplied is consumed in heating the gas. Therefore, a steady state cannot be
reached and the machine exceeds its temperature/pressure limits after some time. To
achieve steady-state operation, cooling is required as soon as the maximum volume
is reached, see Fig. 11.6b. This is consistent with Planck’s formulation of the second
law of thermodynamics, which states that cooling is required for the continuous
conversion of heat into mechanical work.

Theorem 15.2 At least two reservoirs are therefore required: a hot reservoir from
which heat can be taken and a cold reservoir to which heat can be released.
358 15 Second Law of Thermodynamics

Fig. 15.3 Thermal engine (open system inside) in black box notation

15.1.1 The Thermal Engine

Figure 15.3 shows a so-called thermal engine4 in black box notation, i.e. a technical
machine that converts thermal energy into mechanical energy: Theorem 15.2 states
that two reservoirs are needed and heat must be released. The thermodynamic proof
is given in this section. Following the first law of thermodynamics in steady state
operation, the amount of internal energy inside the system must not vary with respect
to time, so that dU
dt
= 0, i.e.

dU
= 0 = Q̇ − Q̇ 0 − P (15.4)
dt

According to Fig. 15.3a, Q̇ increases the internal energy and thus have a positive
sign, while Q̇ 0 and P leave the system and have a negative sign5 in Eq. 15.4. This
notation corresponds to the stationary principle that incoming and outgoing energy
are identical:
Q̇ = Q̇ 0 + P. (15.5)

The question is how much power P can be gained from the thermal energy Q, i.e.

P = Q̇ − Q̇ 0 . (15.6)

Consequently, the cooling Q̇ 0 reduces the maximum output power and should there-
fore be as small as possible. However, the first law of thermodynamics reaches its

4 Anyhow, this cycle is designed as an open system inside, i.e. the required technical components
are flowed through by a mass flow.
5 Nevertheless, positive absolute values are always used in the equation.
15.1 Formulation According to Planck—Clockwise Cycle Processes 359

limits at this stage, as it leads to one equation with two unknowns.6 In order to quan-
tify the maximum power output, an entropy balance is formulated: Entropy is a state
value, so its amount must be constant over time within the machine in steady-state
operation. The entropy balance therefore obeys

dS Q̇ Q̇ 0
=0= − + Ṡi (15.7)
dt T T0

However, in other words, the incoming entropy7 balances the outgoing entropy, see
also Fig. 15.3b:
Q̇ Q̇ 0
+ Ṡi = . (15.8)
T T0

To calculate the entropy carried along with the heat, the temperatures T respectively
T0 within the system at which the heat exceeds the system boundary are needed. In
this case, following the temperature gradient, it is

T > T0 . (15.9)

Obviously, cooling by Q̇ 0 is required in order to release entropy. Without heat release,


the entropy balance cannot be satisfied. The cooling demand can be easily calculated
and is  

Q̇ 0 = T0 + Ṡi . (15.10)
T

The more entropy is produced within the machine, the more cooling is required.
Furthermore, the greater the temperature of the cold reservoir, the more heat needs
to be released. Since the cooling demand is now known, the power output can be
calculated with Eq. 15.6:  

P = Q̇ − T0 + Ṡi . (15.11)
T

In a different notation it is
 
T0
P = Q̇ 1 − − Ṡi T0 . (15.12)
T

The efficiency of this machine is defined with the so-called thermal efficiency:

Benefit
ηth = . (15.13)
Effort

6 The equation is underdetermined.


7 Note that generation of entropy is a source term, so it is going into the system.
360 15 Second Law of Thermodynamics

For this case it obeys


P
ηth = . (15.14)

It follows, that
 
T0 T0
ηth = 1 − − Ṡi (15.15)
T Q̇

It is obvious that the best machine, i.e. the machine with the highest possible effi-
ciency, has no internal entropy generation, Ṡi = 0. Such a machine is denoted Carnot
machine with an efficiency of
T0
ηth, C = 1 − . (15.16)
T
Obviously, the thermal engine (TE) works in-between a hot and cold reservoir. Due
to the temperature gradient, according to the second law of thermodynamics, there
is a heat flux that follows the temperature gradient between the two reservoirs, see
Chap. 14. A thermal engine takes a heat flux from the hot reservoir and converts
it partly into mechanical energy, while releasing some of the energy to the cold
reservoir.

15.1.2 Why Clockwise Cycle?

So far, the thermal engine, which converts heat into work, has been represented in
a black-box notation. This section clarifies how a thermodynamic process within
the machine can be characterised. A distinction is made between cyclic processes
that use piston-engines, i.e. closed systems, and those that are designed as fluid-flow
machines, i.e. open systems.

Realised with closed systems

The process within the machine may consist of four successive changes of state, see
Fig. 15.4. Since the machine is to be operated under steady-state conditions, the fluid
must follow a thermodynamic cycle. As the fluid is in a closed system the volume
work can be represented in a p, V -diagram. For simplicity, the process shown in this
figure is intended to be reversible and outer energies are ignored, i.e.

W = WV . (15.17)

According to the p, V -diagram work is released by the fluid during expansion


(3)→(4), i.e.
15.1 Formulation According to Planck—Clockwise Cycle Processes 361

Fig. 15.4 Thermal engine (closed system) in a p, V -diagram

4
W34 = − p dV < 0. (15.18)
3

For the compression (1)→(2) work is supplied, i.e.

2
W12 = − p dV > 0. (15.19)
1

Work for compression respectively work that is released is represented by the areas
under the changes of state in a p, V -diagram. In order for the overall process to
release work, i.e. the benefit of a thermal engine, the change of state for expansion
must be at a greater pressure level than the change of state for compression. According
to Fig. 15.4 this only works for a clockwise cycle, which releases more work than it
consumes. The effective work Weff is represented by the enclosed area of the changes
of state in a p, V -diagram for a reversible process neglecting the outer energies.
However, in black box notation, the amounts of heat Q in , Q out and work W are
usually represented for closed cycle processes. In order to determine the power output
P of the machine, the speed n and the revolutions per working stroke8 K must be
taken into account, i.e.
n n
P = Weff = mweff . (15.20)
K K

8 Four-stroke K = 2, two-stroke K = 1
362 15 Second Law of Thermodynamics

Fig. 15.5 Thermal engine (open system) in a p, v-diagram

Realised with open systems

Cyclic processes based on fluid flow machines, i.e. consisting of components that
are passed by a mass flux ṁ, are exemplified in Fig. 15.5. For simplicity, the process
shown in this figure is intended to be reversible and outer energies are ignored, i.e.

wt = y. (15.21)

The specific technical work in each component is purely composed of specific pres-
sure work.
According to the p, v-diagram specific technical work is released by the fluid
during expansion (3)→(4), i.e.

4
wt,34 = y34 = v d p < 0. (15.22)
3

For the compression (1)→(2) specific technical work is supplied, i.e.

2
wt,12 = y12 = v d p > 0. (15.23)
1
15.1 Formulation According to Planck—Clockwise Cycle Processes 363

Specific technical work for compression respectively specific technical work that is
released is represented by the areas projected from the change of state to the p-axis
in a p, v-diagram. In order for the overall process to release specific technical work,
i.e. the benefit of a thermal engine, the change of state for expansion must be at
a greater specific volume level than the change of state for compression. Accord-
ing to Fig. 15.5 this only works for a clockwise cycle, which releases more specific
technical work than it consumes. The specific effective work weff is represented by
the enclosed area of the changes of state in a p, v-diagram for a reversible process
neglecting the outer energies.
However, in black box notation, the heat fluxes Q̇, Q̇ 0 and power P are usually
represented for open-system cycle processes. The power can be easily calculated
with the mass flux ṁ, i.e.
P = ṁ wt . (15.24)

15.2 Formulation According to


Clausius—Counterclockwise Cycle processes

According to Clausius heat can not be lifted from a cold reservoir to a hot reservoir
without any external efforts, cf. Fig. 15.6. If such a machine existed, the first law of
thermodynamics, see Fig. 15.6a, would obey

Q in = Q out . (15.25)

In steady state the energy being supplied to the system, needs to be equalised by the
released energy. Thus, the first law of thermodynamics does not show any contra-
dictions. As mentioned previously, the first law of thermodynamics does not restrict

Fig. 15.6 Contradiction to the second law according to Clausius


364 15 Second Law of Thermodynamics

energy conversion. In order to evaluate the feasibility of such a process, a balance


of entropy, see Fig. 15.6b needs to be applied. Again, in steady state operation, the
entropy inside the machine needs to be constant over time, so that released entropy
is balanced by supplied entropy:
Q in Q out
+ Si = . (15.26)
T0 T
     
IN OUT

This, in combination with the first law of thermodynamics, leads to


Q in Q in
Si = − < 0. (15.27)
T T0

The entropy generation is negative, since, according to the temperature gradient


between hot and cold reservoir, T is greater than T0 , see Fig. 15.6. Consequently,
such a process is impossible, cf. Sect. 13.7.4. The amount of supplied entropy due to
heat and internal generation of entropy is greater than the amount of entropy that is
released by the heat. This results in the second law of thermodynamics according to
Clausius:
Theorem 15.3 Heat can never pass by itself from a colder to a warmer reservoir.

15.2.1 The Cooling Machine/Heat Pump

Nevertheless, the question arises as to how a machine can be realised that takes
thermal energy from a cold reservoir and transfers it to a warm reservoir. With
respect to Fig. 15.6b, the amount of entropy leaving the machine must be increased
so that the machine can be operated in a steady state. According to the first law of
thermodynamics, the heat respectively the entropy leaving the machine increases in
steady state when additional work is supplied, i.e.

Q in + W = Q out . (15.28)

The formulation with the heat amounts and the required work is usually applied if
the machine would be realised by a closed system, e.g. cylinder/piston. The power
then result from the speed of the machine, cf. Sect. 15.1.2. However, it is
  
Q̇ in dt + P dt = Q̇ out . (15.29)
t t t

If the machine is internally realised by open systems that are passed by a mass flow,
one usually chooses the power formulation,9 see Fig. 15.7a. The power formulation

9 This approach is followed now.


15.2 Formulation According to Clausius—Counterclockwise Cycle processes 365

Fig. 15.7 Cooling machine/heat pump in black box notation

is the derivative of Eq. 15.29 with respect to time, i.e.

Q̇ in + P = Q̇ out . (15.30)

The internal energy of the system therefore remains constant, i.e.

dU
= Q̇ in + P − Q̇ out = 0. (15.31)
dt

However, the application of the first law of thermodynamics leads to two unknowns,
namely P and Q̇ out , so that further information is required. This can be obtained by
an entropy balance in steady state:

dS Q̇ in Q̇ out
=0= + Ṡi − (15.32)
dt T0 T

respectively, in an entropy-in-equals-entropy-out notation

Q̇ in Q̇ out
+ Ṡi = . (15.33)
T0 T

Thus, the required heat to be released obeys

T
Q̇ out = Q̇ in + Ṡi T. (15.34)
T0

According to the first law of thermodynamics the required power for the operation
then is
P = Q̇ out − Q̇ in . (15.35)
366 15 Second Law of Thermodynamics

Substituting the released heat results in


 
T
P = Q̇ in − 1 + Ṡi T. (15.36)
T0

The conclusion is, that


• the more entropy is generated within the machine, the more power P is required
and
• the greater the temperature difference between cold and hot reservoir is, the more
power P is required.
The efficiency of the machine is

Benefit
ε= . (15.37)
Effort
Usually the efficiency of a machine is indicated with the Greek letter η. However,
the efficiency η by definition is in the range of 0 . . . 1. Since a heat pump, following
Eq. 15.37, has an efficiency greater than 1, the coefficient of performance ε is used
instead of an efficiency η, though its physical interpretation remains the same. A
distinction between cooling machine (refrigerator) and heat pump follows in the
next section.
The temperature profile in Fig. 15.7 needs to be clarified: Theorem 14.2 has shown
that heat always follows a temperature gradient, i.e. in order to transfer heat at the cold
end into the machine, a temperature difference is required, so that a realistic profile
has the course shown in Fig. 15.8. Figure 15.7 merely indicates that the machine
operates within a temperature potential.

Fig. 15.8 Temperature


gradient clarification
15.2 Formulation According to Clausius—Counterclockwise Cycle processes 367

Cooling machine

In case the machine is operated as a cooling machine, the purpose is to take heat
from a cold reservoir.10 Objects in the refrigerator, for example, release heat at a
low temperature level. This heat is supplied to the cooling machine.11 Consequently,
according to Eq. 15.37 the coefficient of performance for a cooling machine (CM) is

Q̇ in
εCM = (15.38)
P

Substitution of P leads to

Q̇ in
εCM =
. (15.39)
Q̇ in T
T0
− 1 + Ṡi T

The best machine works reversibly,12 i.e. Ṡi = 0, so that

T0
εCM,C = (15.40)
T − T0

Heat pump

In case the machine is operated as a heat pump, the purpose is to release heat to
the hot reservoir.13 However, the technical design of a heat pump is identical with
a cooling machine. The coefficient of performance, see Eq. 15.37, for a heat pump
(HP) obeys
Q̇ out
εHP = (15.41)
P

Dividing the first law of thermodynamics

Q̇ out = Q̇ in + P (15.42)

by the power P results in


Q̇ out Q̇ in
= + 1. (15.43)
P P

10 This machine is designed for cooling.


11 How this is realised is explained in Part II.
12 This is a Carnot machine.
13 This machine is designed for heating.
368 15 Second Law of Thermodynamics

Fig. 15.9 Cooling machine/heat pump (closed system) in a p, V -diagram

The introduction of the power coefficients leads to

εHP = εCM + 1 (15.44)

Hence, it is
Q̇ in
εHP =
+ 1. (15.45)
Q̇ in T
T0
− 1 + Ṡi T

Again, the best machine works reversibly,14 i.e. Ṡi = 0, so that:

T
εHP,C = (15.46)
T − T0

15.2.2 Why Counterclockwise Cycle?

The explanation is the same as in Sect. 15.1.2. In this case the machine needs to
be supplied with work, see Fig. 15.9. For visualisation the thermodynamic cycle is
shown in a p, V -diagram and may be realised as a closed cylinder/piston system. To
keep it simple, the process shown in this figure shall be reversible and outer energies
are ignored, i.e. W = WV .
According to the p, V -diagram work is released by the fluid during expansion
(2)→(1), i.e.

14 This is a Carnot machine as well.


15.2 Formulation According to Clausius—Counterclockwise Cycle processes 369

1
W21 = − p dV < 0. (15.47)
2

For the compression (1)→(2) work is supplied, i.e.

2
W12 = − p dV > 0. (15.48)
1

Work for compression respectively work, that is released, are both represented by
the areas beneath the changes of state in a p, V -diagram. In order to achieve, that the
overall process is supplied with work,15 the change of state for expansion requires
to be on a lower pressure level than the change of state for compression. According
to Fig. 15.9, this only works with a counterclockwise cycle, which always consumes
more work than it releases. The effective work is represented by the enclosed area
of the changes of state in a p, V -diagram for a reversible process with neglect of the
outer energies.

Problem 15.1 A heat pump uses ground water with a temperature of 8 ◦ C as a heat
source to heat a room with a temperature of 40 ◦ C. The required thermal power for
the heating purpose is 50 kW. The maximum electrical power consumption shall be
10 kW. Is the given task feasible? What is the heat flux that is extracted from the
ground water?

Solution

According to Fig. 15.10a the first law of thermodynamics reads as

Q̇ in + P = Q̇ out . (15.49)

The required heat from the ground water, i.e. the cold reservoir is

Q̇ in = Q̇ out − P = 40 kW. (15.50)

However, this does not answer the question, whether such a process is possible. Thus,
a balance of entropy, cf. Fig. 15.10b, is applied, i.e.

Q̇ in Q̇ out
+ Ṡi = . (15.51)
T0 T

15 This is the only way to lift heat against the temperature gradient, as the entropy balance has

shown.
370 15 Second Law of Thermodynamics

Fig. 15.10 Solution to Problem 15.1

The entropy generation results in

Q̇ out Q̇ in
Ṡi = − . (15.52)
T T0

This leads to
kW
Ṡi = 0.017395 > 0. (15.53)
K

Since Ṡi > 0 the process is irreversible but possible, cf. Sec. 13.7.4.

Problem 15.2 A specific work of w = 100 kg kJ


is supplied to a thermodynamic cycle.
The machine releases a specific heat of q0 = 160 kgkJ
at an ambient temperature of
Tenv = 290 K.
(a) Is the thermodynamic cycle a thermal engine, a heat pump or a cooling machine?
(b) Which key figure characterises this cyclical process and what value does it reach?
(c) The thermodynamic cycle shall be reversible and the heat absorption takes place
at constant temperature. What is the temperature of the cold reservoir?

Solution

(a) Since the machine is effectively consuming work, it must be a counterclockwise


cycle. The heat is released at ambient temperature, so it is most likely a cooling
machine.
(b) A counterclockwise cycle is characterised by the coefficient of performance. In
this case it reads as
Benefit q
εCM = = . (15.54)
Effort w
15.2 Formulation According to Clausius—Counterclockwise Cycle processes 371

Applying the first law of thermodynamics results in

q + w = q0 . (15.55)

Hence, it is
kJ
q = q0 − w = 60 . (15.56)
kg

The coefficient of performance follows accordingly


q
εCM = = 0.6. (15.57)
w
(c) The balance of entropy for that case obeys
q q0
+ si = . (15.58)
T Tenv

Since the machine works reversibly, there is no generation of entropy, i.e. si = 0.


Thus, it is
q
T = Tenv = 108.75 K. (15.59)
q0

15.3 The Carnot-Machine

The recently discussed thermodynamic cycles have so far been presented as black
boxes. Clockwise cycles, i.e. thermal engines, convert heat into work. Counterclock-
wise cycles are heat pumps or refrigeration machines and lift heat from a lower tem-
perature level to a higher temperature level. The highest efficiency of such machines
can be achieved if they are operated reversibly. In this section, such a reversible
machine, named after Nicolas Leonard Sadi Carnot (1796–1832), is studied. The
Carnot process shows how heat at a given temperature can be most efficiently con-
verted into work by a thermal engine. It also illustrates how heat can most efficiently
pass from a cold to a hot reservoir in terms of a refrigerating machine or heat pump
respectively.
The idealised Carnot cycle consists of two reversible isothermal and two reversible
adiabatic, i.e. isentropic, changes of state. Every change of state is reversible—
including the heat transfer out of and into the reservoirs. In order to achieve this
reversible heat transfer, the heat transfer into respectively out of the machine must
take place at the same temperature as the reservoirs, i.e. with a temperature potential
T → 0, see Chap. 14. Any temperature difference generates entropy during heat
transfer and is therefore irreversible. It has been shown that the greater the temperature
difference, the more entropy is generated. Nevertheless, the heat transfer follows:
372 15 Second Law of Thermodynamics

Q̇ = k A (T − Tenv ) . (15.60)
  
T

The heat flux is therefore proportional to the temperature difference and the surface
on which the heat transfer takes place. Consequently, for a reversible heat transfer
( Ṡi → 0 at T → 0) it follows that

k A → ∞. (15.61)

Accordingly, the theoretical Carnot machine must be infinitely large.

15.3.1 The Carnot-Machine—Clockwise Cycle

First, the focus is on a clockwise Carnot machine, i.e. a thermal engine,16 see
Fig. 15.11a. In this figure, the machine is shown as a black box. It is important to
emphasise that the heat transfers between the machine and the two reservoirs occur
at T = 0 so that the entire process remains reversible.17 The process does not inter-
nally generate entropy, i.e. si = 0, since the Carnot machine follows a perfect, i.e.
reversible process. Consequently, there is neither dissipation nor other imperfections.
It is now examined how the process runs internally, see also Fig. 15.11b, c, which
additionally show a p, v- as well as a T, s-diagram. Obviously, the changes of state
in a Carnot machine are as follows:
• (1) → (2): Reversible, isothermal expansion
Heating is required to keep the temperature constant during expansion. This is
visible in the T, s-diagram, since the specific entropy rises. Since there is no dissi-
pation, the specific heat supplied causes the specific entropy to increase. Following
the p, v-diagram, the specific volume rises while the pressure decreases. Specific
work is released by the machine.
• (2) → (3): Reversible, adiabatic expansion
During this isentropic change of state, the specific entropy remains constant. The
temperature drop is due to the work released during the expansion.
• (3) → (4): Reversible, isothermal compression
Cooling is required to keep the temperature constant when compressing the fluid.
This is visible in the T, s-diagram as the specific entropy decreases. Since there
is no dissipation, the specific heat released is the cause of the decreasing spe-
cific entropy. Following the p, v-diagram, the specific volume decreases while the
pressure rises. Specific work is supplied to the machine.

16The machine is to be constructed exemplarily as a closed system.


17This refers to an extended system boundary that includes the heat transfer between the machine
and reservoirs.
15.3 The Carnot-Machine 373

Fig. 15.11 Clockwise Carnot process—thermal engine

• (4) → (1): Reversible, adiabatic compression


During this isentropic change of state, the specific entropy stays constant. The
temperature rise is due to the supplied work during the compression.
Obviously, from (1)→ (2)→ (3) specific work is released and can be illustrated by
the area beneath18 the change of state in a p, v-diagram. The supplied specific work
is given by the area beneath (3)→ (4)→ (1). Thus, the effective specific work is the
enclosed area in the p, v-diagram. The specific work gained follows from the first
law of thermodynamics19 for thermodynamic cycles, see Sect. 11.2:

q+ w = 0. (15.62)

18 If the Carnot machine is realised by a closed system.


19 This equation means that the absolute value of the incoming energy is equal to the absolute value
of the outgoing energies.
374 15 Second Law of Thermodynamics

Fig. 15.12 Clockwise


Carnot process—
T, s-diagram

Applied to the machine shown in Fig. 15.11 it reads as20

qin = weff + qout . (15.63)

The specific heats qin and qout can be visualised in a T, s-diagram, see Fig. 15.12. As
the entire process, i.e. also the heat transfer, is reversible (ψ = 0), the areas beneath
the changes of state (1)→ (2) and (3)→ (4) show the transferred specific heats.
According to
weff = qin − qin (15.64)

the enclosed hatched area in Fig. 15.12 represents the effective specific work.
The thermal efficiency in this case follows
weff
ηth = . (15.65)
qin

Substitution of the effective specific work by Eq. 15.64 results in

qin − qout
ηth = . (15.66)
qin

The specific heats exchanged can be easily calculated using the T, s-diagram. Since
all steps are reversible, the specific heats can be calculated based on the rectangular
areas. The specific heats in Eq. 15.66 need to be positive since the sign has already
been taken into account by the first law of thermodynamics, cf. Eq. 15.63. Thus,
according to the T, s-diagram the specific heats are

qin = Tmax (s2 − s1 ) (15.67)

20As for all systems in a steady state, a balance of entering and leaving energies applies. The
energies in the equation are all used as absolute values. The sign is automatically taken into account
by deciding whether the flow enters or leaves.
15.3 The Carnot-Machine 375

respectively
qout = Tmin (s3 − s4 ) . (15.68)

Since s2 = s3 and s1 = s4 it follows

Tmax (s2 − s1 ) − Tmin (s2 − s1 )


ηth = . (15.69)
Tmax (s2 − s1 )

Rearranging results in
Tmin
ηCarnot = 1 − (15.70)
Tmax

This is exactly the same result that has already been derived with the black box
approach, see Eq. 15.16.

15.3.2 The Carnot-Machine—Counterclockwise Cycle

The Carnot machine21 can also be operated as a counterclockwise cycle, see


Fig. 15.13. A reservoir releases heat Q in at Tmin , which is supplied to a Carnot pro-
cess. The machine releases a greater amount of heat at a higher temperature level
Tmax . Heat is thus lifted against the temperature gradient from a cold to a hot reser-
voir. To satisfy the steady-state entropy balance, the heat released must be greater
than the heat supplied. In special notation, this results in

qout > qin . (15.71)

According to the first law of thermodynamics this can be achieved by supplying


specific work
qout = qin + weff . (15.72)

Following the principles of the clockwise Carnot cycle discussed previously, the
specific heats can be easily calculated, i.e.

qin = Tmin (s1 − s4 ) (15.73)

respectively
qout = Tmax (s2 − s3 ) . (15.74)

Hence, the coefficients of performance can be calculated as follows:

21 The machine is to be constructed exemplarily as a closed system.


376 15 Second Law of Thermodynamics

Fig. 15.13 Counterclockwise Carnot process—cooling machine/heat pump

• Cooling machine
qin qin
εCM = = (15.75)
weff qout − qin

Substitution of the specific heats results in

Tmin (s1 − s4 )
εCM = . (15.76)
Tmax (s2 − s3 ) − Tmin (s1 − s4 )

Hence, with s3 = s4 and s1 = s2 it finally reads as

Tmin
εCM = (15.77)
Tmax − Tmin

• Heat pump
qout qout
εHP = = (15.78)
weff qout − qin
15.3 The Carnot-Machine 377

Substitution of the specific heats results in

Tmax (s2 − s3 )
εHP = . (15.79)
Tmax (s2 − s3 ) − Tmin (s1 − s4 )

Hence, with s3 = s4 and s1 = s2 it finally reads as

Tmax
εHP = (15.80)
Tmax − Tmin

Both equations have already been derived using the black box approach, see
Eqs. 15.38 and 15.46.
Chapter 16
Exergy

So far, the law of conservation of energy, i.e. the first law of thermodynamics, has been
discussed, which quantifies the changeability of energy. However, this changeability
is limited, so that thermal energy cannot be completely converted into mechanical
energy in a stationary process. The second law of thermodynamics can be applied
to evaluate the limitations of energy conversion. In Chap. 15 a clockwise Carnot
machine has been presented: A machine operating between two thermal reservoirs
to convert thermal energy into maximum mechanical work. The best efficiency is
determined by the minimum and maximum temperatures in between the machine
operates. This principle is important to understand the thermodynamic idea of exergy
as the maximum working capability of any form of energy. The significance of exergy
is presented in this chapter.
To get an idea of what exergy is all about, Fig. 16.1 provides a good introduction to
the subject: Think of two identical tea cups, each filled with the same amount of tea at
the same temperature, e.g. 50 ◦ C, and the same pressure, so that their thermodynamic
states are identical. Consequently, both have the same content of internal energy, i.e.

U(a) = U(b) . (16.1)

However, the cups are in two different locations: one at the North Pole, i.e. at −20 ◦ C,
and the other in the tropics, i.e. at 50 ◦ C. Now the systems are to be evaluated in
terms of their energy conversion potential, i.e. their working capability. A thermal
engine can be used for such an evaluation. In case 1, see Fig. 16.1a, the tea itself
represents a hot reservoir, while the environment is the cold reservoir. The difference
in temperature between the two reservoirs triggers a heat flux. Therefore, a thermal
engine can be installed to connect the two reservoirs. As a result, the incoming heat
flux can be collected and converted into mechanical work, while part of this energy
is released to the cold reservoir due to the second law of thermodynamics. According
to Eq. 15.15, the conversion rate increases as the temperature difference between the
two reservoirs increases.
© Springer Nature Switzerland AG 2022 379
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_16
380 16 Exergy

Fig. 16.1 Evaluation of thermal energy sources

However, in case 2, cf. Fig. 16.1b, there is no temperature difference, i.e. hot (tea)
and cold ( environment) reservoir are already in a thermodynamic equilibrium. It is
therefore not possible to operate a thermal engine in between. From a thermodynamic
point of view, the tea has no working capability.
The maximum working capability that can be achieved is called exergy. As the
introductory example has shown, the environment, i.e. the cold reservoir, must be
taken into account to evaluate the exergy of systems.
Definition 16.1 Exergy is the maximum work that can be gained from the energy
of a system when it is driven into thermodynamic equilibrium with its environment.
Systems in non-ambient states possess exergy. Systems at ambient pressure and
temperature have no exergy and are worthless for any technical application. Strictly
speaking, exergy is not a state value, since it also depends on the state of the envi-
ronment. According to Definition 16.1, energy E can be split into exergy E x and
an energetically worthless part that cannot be further converted into work, denoted
anergy Bx . Thus, it reads as

E = E x + Bx = const. (16.2)

Unlimited convertible forms of energy consist exclusively of exergy, e.g. electrical


energy or mechanical energy. Non-convertible forms of energy consist exclusively
of anergy, e.g. thermal energy at ambient temperature. Partially convertible energy
consists of exergy and anergy. Unlike energy,1 exergy can be consumed and can thus
decrease.2 Exergy is therefore not a conservation variable, but an extensive variable

1 The law of conservation of energy states that energy must be constant.


2 To keep energy constant, anergy increases when exergy decreases.
16 Exergy 381

Fig. 16.2 Gaining work


from heat

that can be balanced.3 In the following sections, the exergy of different forms of
energy is derived. This is necessary to balance exergy of open and closed systems.
In addition, the cause for the loss of exergy is explained thermodynamically.

16.1 Exergy of Heat

In this first section, the exergy of heat E x,Q is determined, i.e. the maximum amount of
work that can be gained from the thermal energy Q available at a temperature T . For a
heat flux Q̇ there is a corresponding exergy flux Ė x,Q , which is now to be determined.
The basic idea for determining the exergy is to bring the heat into a thermal engine,4
see Fig. 16.2. A thermal engine splits thermal energy into mechanical and thermal
energy. Consequently, a thermal engine can be utilised as an Energy-in-Exergy-and-
Anergy-Decomposition-Machine. The maximum work can only be achieved in a
reversible process, since any irreversibility leads to entropy generation and thus to a
loss of working capability. The efficiency of such a process has already been derived
with Eq. 15.15 in Chap. 15:

P T0 T0 Ṡi
ηth = =1− − (16.3)
Q̇ T Q̇

The requirements for the thermal engine can be derived from this equation:
• Entropy generation:
Ṡi = 0 (16.4)

The introduced energy-in-exergy-and-anergy-decomposition-machine must be


free of any generation of entropy, i.e. Ṡi = 0, in order to achieve the maximum

3 This exergy balancing is usually performed at a constant environment.


4 A thermal engine has been introduced as a clockwise thermodynamic cycle, that converts heat
into work, cf. Chap. 15.
382 16 Exergy

Fig. 16.3 Energy-in-exergy-


and-anergy-decomposition-
machine

work. Consequently, a Carnot machine is the best option to convert heat Q respec-
tively a heat flux Q̇ into work respectively power.
• Heat release:
T0 → Tenv (16.5)

To maximise the efficiency respectively the gained work, the released heat must
be transferred at the lowest possible temperature T0 . Obviously, ambient temper-
ature Tenv is the lower limit of the process, otherwise it would not be possible
to release heat from the machine to the environment.5 In this theoretical limiting
case, no irreversibility occurs during heat transfer because the temperature dif-
ference reaches zero, see Sect. 15.3. As already mentioned, the machine must be
very large to be able to transfer heat with a temperature difference of T → 0. If
the Carnot process releases heat above ambient temperature, working capability is
wasted,6 as the released heat above ambient temperature could be used to operate
another Carnot machine, since there is still a temperature potential with respect to
the environment.
Under these boundary conditions, see also Fig. 16.3, the power output is maximised
and the exergy of the heat follows

!
Ė x,Q = P. (16.6)

The entering heat flux Q̇ at a temperature T > Tenv consists of exergy Ė x,Q and
anergy Ḃx,Q . Due to the second law of thermodynamics a heat flux Q̇ out must be
released. Since the released heat flux is already at ambient temperature Tenv it does
not contain any exergy, i.e. no additional thermal engine can be operated due to
the missing temperature gradient. Hence, Q̇ out is pure anergy. Under the boundary
conditions discussed, the power output P is maximum and pure exergy, since it is
mechanical work. The principle that heat must be released at ambient temperature

5 In case the temperature at the cold end of the machine was below ambient temperature, there
would be a heat flux from environment into the machine.
6 And the efficiency would not be maximum.
16.1 Exergy of Heat 383

Fig. 16.4 Heat at constant temperature → Evaluation of the exergy of the heat q@T with a Carnot-
machine

to achieve maximum power can be explained in more detail with a T, s-diagram,


cf. Fig. 16.4. In this figure a clockwise cycle7 is shown. Specific heat at a constant
temperature T , i.e. q@T is to be evaluated, is supplied to the thermal engine from
(1)→(2). The amount of specific heat is given by the integral of this change of
state. The specific heat released can be represented in a similar way: In this case,
the area under (3)→(4) represents the specific heat released. However, the enclosed
area (1)(2)(3)(4) indicates the output of specific net work. In order to determine
the specific exergy this enclosed area has to be maximised. In Fig. 16.4a specific
exergy is wasted, since the specific work is not yet maximised. Instead, Fig. 16.4b
shows the maximum specific work. The upper limit of the enclosed area must not be
touched, because the specific heat at this specific temperature T has to be investigated.
Consequently, the specific heat release needs to be optimised. The maximum specific
work is achieved, when the specific heat is released at ambient temperature. It is not
possible to fall below ambient temperature since in that case it would be unfeasible
to emit heat due to the second law of thermodynamics.8 Under these conditions the
specific exergy ex,Q can be illustrated as the hatched area (1)(2)(3)(4). Accordingly,
the specific anergy bx,Q follows (3)(4)(a)(b), see Fig. 16.4b. The supplied specific
heat q@T is the enclosed area (1)(2)(b)(a), i.e.

q = ex,Q + bx,Q . (16.7)

Following the thermal efficiency of a Carnot-machine, see Eq. 15.70, the exergy can
be calculated as follows:
Pmax Tenv
ηCarnot = =1− . (16.8)
Q̇ T

7The Energy-in-Exergy-and-Anergy-Decomposition-Machine must be a clockwise cycle.


8The second law of thermodynamics states that heat always follows a temperature gradient from
hot to cold.
384 16 Exergy

Solving for Pmax results in


 
! Tenv
Ė x,Q = Pmax = Q̇ 1 − (16.9)
T
  
=ηCarnot

If not a flux but a amount of heat is of interest, it follows accordingly


 
! Tenv
E x,Q = Wmax = Q 1− (16.10)
T

In order to evaluate the exergy of heat, it is therefore necessary to know the tem-
perature at which the heat is available: the higher the temperature compared to the
ambient temperature, the more exergy the heat possesses.

16.1.1 Heat at Constant Temperature

For the special case of an isothermal expansion, the exergy can easily be represented
in a T, s-diagram, see Fig. 16.5. In case the change of state is reversible, the expla-
nation is exactly as given above: The specific exergy is represented by the hatched
area between T and Tenv , see Fig. 16.5a. The integral of (1) to (2) equals the specific
heat, as there is no dissipation. However, in case (b), i.e. irreversible, isothermal
expansion, the situation is somewhat different, as specific dissipation also occurs.
Since the temperature is constant, the integral form (1) to (2) is divided into a part

Fig. 16.5 Exergy of heat at isothermal expansion according to [3]


16.1 Exergy of Heat 385

Fig. 16.6 Non-isothermal heat supply (1)→(2)

belonging to the specific heat q and another belonging to the specific dissipation
ψ12 : The entire change of specific entropy is due to one part caused by heating, i.e.
sa,12 , and a second part due to dissipation, i.e. si,12 . Since the specific heat q and the
specific power loss ψ occur at the same temperature, which is constant from (1) to
(2), this approach can be applied.9

16.1.2 Heat at Variable Temperature

In case, heat is transferred at a variable temperature, e.g. the heat exchanger in


Fig. 13.21, the exergy illustration in a T, s-diagram is more complex, see Fig. 16.6.
If the change of state (1)→(2) runs reversibly, the supplied specific heat is the area
beneath the change of state. In combination with the specific heat released at ambi-
ent temperature, the specific exergy can be illustrated by the hatched area, as shown
in Fig. 16.6a. This information can not be gained in case the change of state is
irreversible, see Fig. 16.6b. Since specific heat and specific dissipation occur simul-
taneously, a splitting as shown in Fig. 16.5b cannot be done because there is only
one thermodynamic mean temperature for both specific heat and specific dissipation.
Even under these conditions the specific exergy of specific heat can be calculated
with a generic approach.10 Since the temperature is variable a differential approach
is followed. Equation 16.9 can be applied at any incremental, quasi-static change of
state. Within this incremental step, only a small portion of heat δ Q̇ at the current

9 Thus, the thermodynamic mean temperature is constant as well. However, this is not trivial if T
is not constant, cf. Fig. 16.6.
10 It does not matter whether the change of state is reversible or irreversible.
386 16 Exergy

temperature T is transferred, see Fig. 16.6a:


 
Tenv
d Ė x,Q = 1− δ Q̇. (16.11)
T

In order to calculate the entire exergy, an integration is done, i.e.

2 2  
Tenv
Ė x,Q = d Ė x,Q = 1− δ Q̇. (16.12)
T
1 1

A simpilfication leads to
2
δ Q̇
Ė x,Q = Q̇ − Tenv . (16.13)
T
1

With the thermodynamic mean temperature, see Sect. 13.7.3, it results in


Ė x,Q = Q̇ − Tenv . (16.14)
Tm

Hence, the generic approach for calculating the exergy of heat is


 
Tenv
Ė x,Q = Q̇ 1 − (16.15)
Tm

With Eq. 13.155 it reads as


Ė x,Q = Q̇ − Tenv Ṡa . (16.16)

The entropy flux Ṡa can be substituted by a balance of entropy. For an open system,
for example, it is
Ṡa + Ṡi = ṁ (s2 − s1 ) (16.17)

respectively
Ṡa = ṁ (s2 − s1 ) − Ṡi . (16.18)

Therefore the exergy flux of a heat flux finally obeys

Ė x,Q = Q̇ − ṁTenv (s2 − s1 ) + Tenv Ṡi (16.19)

Obviously, this generic approach can be applied for reversible and irreversible
changes of state.
16.1 Exergy of Heat 387

Fig. 16.7 Exergy of heat—influence of the environmental temperature according to [1]

In case an amount of heat is regarded,11 it follows accordingly


 
Tenv
E x,Q = Q 1− (16.20)
Tm

With Eq. 13.168 it reads as


E x,Q = Q − Tenv Sa . (16.21)

The entropy Sa can be substituted by a balance of entropy. For closed systems it


obeys
Sa + Si = S2 − S1 . (16.22)

Thus, the exergy of heat finally is

E x,Q = Q − Tenv (S2 − S1 ) + Tenv Si (16.23)

Equations 16.13 respectively 16.20 show, that


• at large ambient temperature the exergy of heat is small and the anergy is corre-
spondingly high. If ambient temperature is low, this is reversed.
• the greater the temperature T at which heat occurs, the greater the exergy.
This correlation is shown in Fig. 16.7. It is

Q = E x,Q + Bx,Q . (16.24)

For the anergy it follows


Bx,Q = Q − E x,Q . (16.25)

11 For a closed system, for example.


388 16 Exergy

In combination with the exergy of heat these equations simplify to

Tenv
E x,Q = Q − Q (16.26)
T
and
Tenv
Bx,Q = Q . (16.27)
T
Dividing these two equations gives the so-called lever rule, which is shown in
Fig. 16.7:
E x,Q T − Tenv
= (16.28)
Bx,Q Tenv

16.1.3 Sign of the Exergy of Heat

Figure 16.8 shows the consequences of Eq. 16.29 regarding the sign of exergy of heat.
 
Tenv
Ė x,Q = Q̇ 1 − (16.29)
T
  
=ηCarnot

Obviously, the Carnot factor ηCarnot gets negative for temperatures T < Tenv . In this
figure, the focus is on a reservoir with a temperature T , which is marked by a dashed
line as the system boundary. The following cases can occur:
• Case (a): Heat supply, i.e. Q̇ > 0, at T < Tenv
In this case it is

Fig. 16.8 Sign of exergy of heat


16.1 Exergy of Heat 389
 
Tenv
Ė x,Q = Q̇ 1 − < 0. (16.30)
 T
>0   
<0

According to Fig. 16.8a exergy leaves, so that Ė x,Q < 0. Heat and exergy have
opposite signs.
• Case (b): Heat release, i.e. Q̇ < 0, at T < Tenv
In this case it is  
Tenv
Ė x,Q = Q̇ 1 − > 0. (16.31)
 T
<0   
<0

According to Fig. 16.8b exergy is required,12 so that Ė x,Q > 0. Heat and exergy
have opposite signs.
• Case (c): Heat supply, i.e. Q̇ > 0, at T > Tenv
In this case it is  
Tenv
Ė x,Q = Q̇ 1 − > 0. (16.32)
 T
>0   
>0

According to Fig. 16.8c exergy is required, so that Ė x,Q > 0. Heat and exergy have
the same sign.
• Case (d): Heat release, i.e. Q̇ < 0, at T > Tenv
In this case it is  
Tenv
Ė x,Q = Q̇ 1 − < 0. (16.33)
 T
<0   
>0

According to Fig. 16.8d exergy leaves, so that Ė x,Q < 0. Heat and exergy have the
same sign.

16.2 Exergy of Fluid Flows

This section clarifies how much working capability a fluid flow possesses. The max-
imum working capability is called the exergy of fluid flows and is required to per-
form exergy balances with open systems. A fluid flow has a working capability
due to its mechanical energy, i.e. kinetic and potential energy, and due to its ther-
mal energy when it is in thermodynamic disequilibrium with the environment.13 cf.
Fig. 16.9a. In order to achieve the maximum power output, the fluid is brought into

12To lift Q̇ against the temperature gradient.


13In the previous section it has been shown that thermal energy can be used in a thermal engine to
gain work.
390 16 Exergy

Fig. 16.9 Exergy of a fluid flow

a thermodynamic, i.e. mechanical and thermal equilibrium with its environment,


see Fig. 16.9b. This is realised in a machine that is supplied with the fluid in state
(1) and operates in a steady state. The aim is to bring the fluid into equilibrium
with the environment: Hence, the flow leaves the machine at ambient conditions, i.e.
penv , Tenv , cenv = 0, z env = 0. To bring the fluid from (1) to (env), technical work
and heat can be exchanged with the environment. However, the task is to maximise
the technical work, as this represents the exergy of the fluid flow14 (1), i.e.

!
Ė x,S,1 = −Pt,rev . (16.34)

Obviously, the machine needs to work reversibly, since any dissipation reduces
the power output, cf. Sect. 11.3.4. The fluid must therefore be reversibly, i.e. with
ψ1env = 0, transferred from state (1) to ambient state (env). For this change of state
the first law of thermodynamics can be applied15 :

1
2
Q̇ rev + Pt,rev = ṁ h env − h 1 + cenv − c12 + g (z env − z 1 ) . (16.35)
2

Not only should mechanical losses be avoided, but heat transfer must also be
reversible, see Chap. 14. Thus, heat must be transferred with T → 0. Consequently,
the heat transfer surface must be very large. Since the environment is at rest and does
not contain any potential energy, the first law of thermodynamics states that

1 2
Ė x,S,1 = ṁ h 1 − h env + c1 + gz 1 + Q̇ rev . (16.36)
2

14 Due to the sign convention a power output is negative. Thus, in order to get a positive exergy, a
minus sign is applied in the definition.
15 Following the energy-in-is-balanced-by-energy-out principle under steady state conditions.
16.2 Exergy of Fluid Flows 391

Fig. 16.10 Change of state (1)→(env) with p1 > penv , according to [3]

In a next step the reversible heat transfer Q̇ rev needs to be specified. The idea is
illustrated in Fig. 16.10. The change of state is divided into two substeps to achieve
the maximum power output. At first, the change of state shall be reversible and
adiabatic,16 i.e. from (1) to (1 ). If it was not adiabatic, heat would be transferred at a
temperature other than ambient, so that exergy, i.e. working capability, would also be
transferred. In order to avoid this, the first change of state to (1 ) is done until ambient
temperature Tenv is reached. In a second step, heat can then be reversibly transferred
from (1 ) to (env), since system and environment have the same temperature, i.e.
T → 0. Heat can be reversibly transferred in both directions, but this would require
an infinite heat transfer surface. According to the T, s,-diagram in Fig. 16.10, the
transferred heat then is pure anergy, so it is exergetically neutral,17 i.e. no exergy
is supplied by heat from the environment respectively released to the environment.
With this approach, the reversible heat can be easily calculated, i.e.

δqrev
ds = δsa + δsi = δsa = (16.37)
 T
=0

respectively
δqrev = T ds. (16.38)

Since T = Tenv = const. and s1 = s1 it follows for the reversibly transferred heat:

env


qrev = T ds = Tenv senv − s1 . (16.39)
1

16 Reversible and adiabatic means the change of state is isentropic.


17 Any exchange of exergy across the system boundary would distort the exergetic evaluation of
the fluid flow.
392 16 Exergy

The heat flux follows by multiplying with the mass flow rate ṁ and obeys


Q̇ rev = ṁTenv senv − s1 . (16.40)

Thus, the exergy of the flow follows by substituting the heat in Eq. 16.36, i.e.

1
Ė x,S,1 = ṁ h 1 − h env + c12 + gz 1 + Tenv (senv − s1 ) (16.41)
2

This equation shows that both kinetic and potential energy are pure exergy. For the
anergy of a fluid flow it is

Ḃx,S,1 = ṁ [h env − Tenv (senv − s1 )] (16.42)

The sum of exergy and anergy flux is the overall energy flux of state (1), i.e.

1
Ė x,S,1 + Ḃx,S,1 = ṁ h 1 + c12 + gz 1 = Ė 1 . (16.43)
2

However, let us have a closer look at Fig. 16.10, that shows how the machine
works in a p, v- as well as in a T, s-diagram. The T, s-diagram has already been
explained, as it shows the reversible heat at ambient temperature, i.e. pure anergy,
that is required to reach ambient state (env). In this example, it is obvious that the
liquid flow must be heated because the specific entropy increases from (1 ) to (env).
Due to T → 0 this can only be achieved with an infinite heat transfer area. In this
limiting case, the environment reversibly supplies heat to the system. To interpret
the p, v diagram, it is reasonable to apply the partial energy equation for the open
system from (1) to (env), i.e.
⎡ ⎤
⎢env ⎥
⎢ 1
2 ⎥
⎢ ⎥
Ė x,S,1 = −Pt,rev = −ṁ ⎢ v d p + ψ1env + cenv − c12 + g (z env − z 1 )⎥ .
⎢    2 ⎥
⎣1 =0 ⎦
  
=y1,env
(16.44)
Thus, in the p, v-diagram, see Fig. 16.10, the specific pressure work y1,env is shown.
Since the mechanical energy, i.e. pure exergy, is missing, the specific exergy is not
illustrated in that diagram. Note that the specific pressure work for open systems is
the projected area to the p-axis. In this example, the fluid releases specific pressure
work from (1) to (1 ) and from (1 ) to (env). This is due to p1 > penv after the
isentropic change of state from (1) to (1 ).
Another example is shown in Fig. 16.11. In this case, after the isentropic change
of state to (1 ), it is p1 < penv . According to the T, s-diagram under these conditions
16.2 Exergy of Fluid Flows 393

Fig. 16.11 Change of state (1)→(env) with p1 < penv , according to [3]

a reversible, isothermal cooling from (1 ) to (env) is required as the specific entropy
needs to decrease. Again, the specific heat released is pure anergy, as it occurs at
ambient temperature. The p, v diagram implies that from (1) to (1 ) pressure work
is released, while from (1 ) to (env) pressure work must be supplied. Hence, the
resulting, effective specific pressure work is illustrated as grey area.

16.3 Exergy of Closed Systems

It has been shown how to determine the exergy of heat and fluid flows. In this section,
the focus is on closed systems. The maximum amount of work that can be extracted
from a closed system is investigated. This maximum effective work is called the
exergy of a closed system. According to Fig. 16.12, working capability is based
on mechanical energy, i.e. kinetic and potential energy, as the centre of gravity of
a closed system may be in motion, and on a thermodynamic imbalance with the

Fig. 16.12 Exergy of a


closed system
394 16 Exergy

Fig. 16.13 Change of state (1)→(env) with p1 > penv

environment. The gas in a closed system may have a different state (1) than the
environment (env) so that the piston can release work during the transient balancing
process (1)→(env). However, the change of state from (1)→(env) must be reversible
in order to maximise the work. The exergy of a closed system in state (1) is defined
as18 :
!
E x,1 = −Weff,rev = −Wrev − penv m (venv − v1 ) . (16.45)

The effective work19 Weff,rev has been introduced in Sects. 9.2.3 and 11.2.3. It
describes the work that can be utilised in a cylinder/piston system, when operated
under ambient conditions. The environment supports the compression of a gas, while
it reduces the released work at expansion. It is therefore the effective work that is
associated with the definition of exergy, as it involves the environment. The first law
of thermodynamics for a reversible process reads as20
 
1 2
Wrev + Q rev = Uenv − U1 − m c1 + gz 1 . (16.46)
2

Thus, in a first step, the system is conditioned to ambient temperature by an adiabatic,


reversible change of state, see Fig. 16.13. In this step, heat transfer must be avoided
because heat transferred at a temperature different from ambient temperature contains
exergy. In order to achieve the maximum work output, transfer of exergy at this point
does not make sense. Furthermore, the heat transfer is only reversible if T → 0.
Therefore, if heat transfer is required to reach the ambient state (env), this is done in
a second step once ambient temperature is reached. Heat at ambient temperature is
pure anergy. In case p1 > penv the system needs to be heated from (1 ) to (env), see

18 Due to the sign convention work release is negative. Hence, in order to get a positive exergy, a
minus sign is applied in the definition.
19 In Eq. 16.45 v
env denotes the specific volume the fluid has under ambient conditions, i.e. at Tenv
and penv . It does not represent the specific volume of the environment.
20 Note that the (env)-state is in rest.
16.3 Exergy of Closed Systems 395

T, s-diagram in Fig. 16.13. If p1 < penv , see T, s-diagram in Fig. 16.14, the system
needs to be cooled in the second change of state from (1 ) to (env). Note that heat
transfer under these conditions is reversible21 but requires an infinite heat transfer
surface. According to the T, s-diagram the heat results in

Q rev = mTenv (senv − s1 ) . (16.47)

Thus, the exergy of a closed system obeys



1 2
E x,1 = U1 − Uenv + m c + gz 1 + Tenv (senv − s1 ) − penv (venv − v1 )
2 1
(16.48)

The anergy is

Bx,1 = Uenv + m [ penv (venv − v1 ) − Tenv (senv − s1 )] (16.49)

since the sum of Eqs. 16.48 and 16.50 leads to



1 2
E x,1 + Bx,1 = U1 + m c1 + gz 1 = E 1 . (16.50)
2

However, let us have a closer look at Fig. 16.13, that shows the changes of state in
p, v- as well as in a T, s-diagram. The T, s-diagram has already been explained, as
it shows the reversible heat at ambient temperature, i.e. pure anergy, that is required
to reach ambient state (env). Obviously, in case p1 > penv , the fluid flow has to be
heated, since the specific entropy from (1 ) to (env) increases. Due to T → 0 this
can only be achieved with an infinite heat transfer surface. To interpret the p, v-
diagram, it is reasonable to apply the partial energy equation for a closed system
from (1) to (env), i.e.

env
1
2
Wrev = WV + 
 +Wmech = − p dV + m cenv − c12 + g (z env − z 1 ) .
2
=0 1
(16.51)
Thus, the effective work is22

Weff,rev = Wrev + penv m (venv − v1 ) (16.52)

respectively

21 No entropy is generated, since T = 0 for the heat transfer.


22 See Eq. 9.42.
396 16 Exergy

env
1
2
Weff,rev = − p dV + penv m (venv − v1 ) +m c − c1 + g (z env − z 1 ) .
2
2 env
1
  
WV,eff
(16.53)
Consequently, in the p, v-diagram, see Fig. 16.13, the specific effective volume work
wV,eff is shown. Since the mechanical energy, i.e. pure exergy, is missing, the specific
exergy is not illustrated in that diagram. Note that the specific effective volume work
for closed systems is the projected area to the v-axis reduced by the work at the
environment, see also Fig. 9.7. In this example, the fluid releases specific volume
work from (1) to (1 ) and from (1 ) to (env), since the fluids expands. This is due to
p1 > penv after the isentropic change of state from (1) to (1 ).
Another example is shown in Fig. 16.14. In this case after the isentropic change
of state to (1 ), it is p1 < penv . According to the T, s-diagram under these conditions
a reversible, isothermal cooling from (1 ) to (env) is required as the specific entropy
needs to decrease. Again, the released specific heat is pure anergy since it occurs at
ambient temperature. The specific effective volume work wV,eff is shown as a grey
area in the p, v diagram for this case. Obviously, the specific effective volume work
is composed as follows:
• Let us first consider the change of state from (1) to (1 ) to (c)

From (1) to (1 ) volume work is released (-), i.e. 1, 1 , f,b,1, which is reduced
effectively by the environment (+), i.e. b,d,e,f,b. From (1 ) to (c) volume work is
supplied (+), i.e. 1 , f,b,c,1 , which is supported effectively by the environment (-),
i.e. b,d,e,f,b. Thus, in total the environmental work b,d,e,f,b is neutralised. The
specific effective volume work for this first step is 1, 1 , c,1.

• In the next step the final change of state from (c) to (env) is investigated

From (c) to (env) volume work is supplied (+), i.e. c,env,a,b,c, which is supported
effectively by the environment (-), i.e. a,b,d,env,a. According to the sketch in

Fig. 16.14 Change of state (1)→(env) with p1 < penv


16.3 Exergy of Closed Systems 397

Fig. 16.14 the absolute value of the environmental work is greater than the supplied
work for the compression. Consequently, the overall work that is released (-) from
(c) to (env) results in d,c,env,d.
Problem 16.1 Two identical vertical pipes with a diameter of D = 0.1 m are con-
nected by a thin tube and a valve, see Fig. 2.8. One pipe is filled with water up to a
height of H = 10 m. Water has a density of ρ = 1000 mkg3 , the other tube is initially
empty. Now the valve is opened. After some time, an equilibrium of the water quan-
tity is reached. Calculate the exergy of the system in state (1) respect. in state (2)!
How much exergy is lost during the change of state?
• Specific heat capacity (water): c p = 4.18 kgkJK
• Ambient state ( penv = 1 bar, Tenv = 293.15 K)
• Initial temperature of the water T1 = 293.15 K
• The system does not exchange any energy with the environment

Solution

According to Eq. 16.48 the exergy in state (1) is



1 2
E x,1 = U1 − Uenv + m c1 + gz 1 + Tenv (senv − s1 ) − penv (venv − v1 ) . (16.54)
2

It simplifies as follows, since state (1) is equal with state (env)


• Internal energy
U1 − Uenv = mc p (T1 − Tenv ) = 0 (16.55)

• Outer energies
1 2 H
c1 + gz 1 = g (16.56)
2 2

• Entropy23
Tenv
Tenv (senv − s1 ) = Tenv c p ln =0 (16.57)
T1

• Ambient work24
penv (venv − v1 ) = 0 (16.58)

Thus, the exergy in state (1) results in

H
E x,1 = mg = 3852.4 J. (16.59)
2

23 Applying the caloric equation for incompressible liquids.


24 Water is treated as an incompressible liquid, so that venv = v1 .
398 16 Exergy

The working capability consists only of the potential energy of the water, since the
internal state is already in equilibrium with the environment.

Now, let us investigate state (2). The exergy follows



1 2
E x,2 = U2 − Uenv + m c + gz 2 + Tenv (senv − s2 ) − penv (venv − v2 ) . (16.60)
2 c

It can be simplified, since state (2) is also in rest and the fluid is still incompressible,
i.e. H Tenv
E x,2 = mc p (T2 − Tenv ) + mg + mTenv c p ln . (16.61)
4 T2

The temperature rise has already been calculated in Problem 12.3, i.e. T2 =
293.155867 K. Hence, the exergy in state (2) is

E x,2 = 1926.2 J. (16.62)

In state (2), the exergy is composed of both mechanical and thermal energy, as the
temperature of the system is slightly above ambient temperature. Since the system
does not exchange energy with the environment,25 there is an exergy loss, i.e.

E x,V = E x,1 − E x,2 = 1926.2 J. (16.63)

16.4 Loss of Exergy

Problem 16.1 has shown that a system can lose exergy. Obviously, there must be an
internal reason for this exergy loss, since the discussed system has not transferred any
exergy across its boundary. Although exergy is not a conservation quantity like mass,
momentum and energy, it can be balanced. In this section, the cause of an exergy loss
is investigated. For this purpose, a distinction is made between closed/open systems
and thermodynamic cycles.

16.4.1 Closed System

The loss of exergy E x,V in a closed system can be easily determined by comparing
states (1) and (2). Starting from state (1) exergy increases by supply of exergy, i.e. due
to heat and effective work. Work is pure exergy, while heat contains partial exergy,
see Sect. 16.1. It is postulated that the exergy is reduced by a loss of exergy, as can

25The system is adiabatic, so there is no exchange of exergy of heat. Furthermore, no electrical or


mechanical work, which would be pure exergy, is transferred.
16.4 Loss of Exerg 399

Fig. 16.15 Closed system:


balance of exergy

be seen in Problem 16.1. Thus, according to Fig. 16.15, for state (2) it is

E x,2 = E x,1 + Weff + E x,Q − E x,V . (16.64)

In other words, the exergy in state (2) is equal to the exergy initially available in
state (1) plus the exergy supplied by heat and work from (1) to (2) minus the exergy
loss.26 Consequently, the loss of exergy follows

E x,V = E x,1 − E x,2 + Weff + E x,Q (16.65)

The exergies for a closed system E x,1 as well as E x,2 can be replaced by the corre-
lations derived in Sect. 16.3. This results in

E x,1 − E x,2 = U1 −U2 +


 
1 2
+m c1 − c22 + g (z 1 − z 2 ) + Tenv (s2 − s1 ) − penv (v2 − v1 ) .
2
(16.66)

The effective work reads as27 :

Weff = W12 + penv (V2 − V1 ) . (16.67)

Furthermore, the exergy of heat, cf. Sect. 16.1, reads as

E x,Q = Q 12 − Tenv (S2 − S1 ) + Tenv Si,12 . (16.68)

26
Obviously, the loss of exergy means a sink.
27
Note that W12 is the work that passes the system boundary of the fluid inside the cylinder, whereas
Weff is the work that can effectively be utilised at the piston.
400 16 Exergy

Applying the first law of thermodynamics, i.e.



1
2
W12 + Q 12 = U2 − U1 + m c − c12 + g (z 2 − z 1 ) (16.69)
2 2

results in the loss of exergy for a closed system:

E x,V = Tenv Si,12 (16.70)

Si,12 is the generated entropy caused by friction, mixing and balancing processes
due to temperature-, pressure- or concentration gradients, see Chap. 14. Obviously,
there is a correlation between dissipation, entropy generation and the loss of working
capability, i.e.
12 → Si,12 → E x,V (16.71)

Dissipation should be avoided since it leads to entropy generation and finally to a


loss of working capability, i.e. loss of exergy.
Overview closed system
Figure 16.16 gives an overview of how closed systems are treated. It summarises
what has been investigated so far:

• First law of thermodynamics, see Fig. 16.16a



1
2
Q 12 + W12 = U2 − U1 + m c − c1 + g (z 2 − z 1 )
2
(16.72)
2 2
  
=E a

Fig. 16.16 Overview closed system: balance of energy (a), entropy (b) and exergy (c)
16.4 Loss of Exerg 401

Dissipation does not appear in the first law of thermodynamics: Energy is a con-
servation quantity. There is neither a source nor a sink inside. Energy is not gen-
erated or destroyed. The amount of energy inside can only change by supplying
or releasing energy across the system boundary. Dissipation inside is merely a
transformation of energy driven by the work supplied or the mechanical energies
inside.
– Partial energy equation

W12 = W12,V + W12,mech +12 (16.73)


  
=E a

– Effective work
Weff = W12 + penv (V2 − V1 ) (16.74)

• Second law of thermodynamics, see Fig. 16.16b

S2 − S1 = m · (s2 − s1 ) = Sa,12 + Si,12 (16.75)

Unlike energy, entropy is not a conservation quantity: entropy can be generated


internally and must be accounted for by a source term Si,12 .
• Balance of exergy, see Fig. 16.16c

E x,2 = E x,1 + Weff + E x,Q − E x,V (16.76)

with
E x,V = Tenv Si,12 (16.77)

Unlike energy, exergy is not a conservation quantity: exergy can be destroyed


within a system and must be accounted for by a sink term E x,V .

16.4.2 Open System in Steady State Operation

The loss of exergy flux  Ė x,V in an open system can be easily determined by com-
paring inlet state (1) and outlet state (2). In steady state operation the state of a system
remains constant by time. Further, assuming a constant environment, the amount of
exergy inside the system needs to be constant by time as well. Thus, starting from
state (1), the exergy increases by the supply of exergy, i.e. by heat and technical
power Pt . Note that technical power is pure exergy. It is postulated, that exergy is
reduced by a loss of exergy as seen in Problem 16.1. Thus, according to Fig. 16.17,
state (2) follows28

28 In this figure, the exergy is shown as bubbles that can be counted.


402 16 Exergy

Fig. 16.17 Open system in


steady state: balance of
exergy

Ė x,S,2 = Ė x,S,1 + Pt + Ė x,Q −  Ė x,V . (16.78)

In other words, the exergy in state (2) is equal to the exergy initially available in
state (1) plus the exergy supplied by heat and work from (1) to (2) minus the exergy
loss.29 Rearranging this equation leads to

Ė x,S,1 + Pt + Ė x,Q = Ė x,S,2 +  Ė x,V (16.79)

Theorem 16.1 Thus, the incoming exergy flux is equal to the outgoing exergy flux
for steady-state operation, as the amount of exergy in the system remains constant.
The loss of exergy obeys

 Ė x,V = Ė x,S,1 − Ė x,S,2 + Pt + Ė x,Q (16.80)

The exergies for fluid flows Ė x,S,1 as well as Ė x,S,1 can be substituted by the corre-
lations that have been deduced in Sect. 16.2. This leads to

1
2
Ė x,S,1 − Ė x,S,2 = ṁ h 1 − h 2 + c1 − c22 + g (z 1 − z 2 ) + Tenv (s2 − s1 ) .
2
(16.81)
The exergy flux due to heat flux obeys Eq. 16.19, i.e.

Ė x,Q = Q̇ 12 − ṁTenv (s2 − s1 ) + Tenv Ṡi . (16.82)

Applying the first law of thermodynamics, i.e.



1
2
Pt + Q̇ 12 = ṁ h 2 − h 1 + c2 − c12 + g (z 2 − z 1 ) (16.83)
2

results in the loss of exergy for an open system in steady state:

29 Exergy loss reduces the exergy, so it is an outgoing flux. It is treated as a sink.


16.4 Loss of Exerg 403

 Ė x,V = Tenv Ṡi (16.84)

Ṡi is the generated entropy caused by friction, mixing and balancing processes due
to temperature-, pressure- or concentration gradients, see Chap. 14. Obviously, there
is a correlation between dissipation, entropy generation and the loss of working
capability, i.e.
˙ → Ṡi →  Ė x,V (16.85)

Dissipation should be avoided since it leads to entropy generation and finally to a


loss of working capability, i.e. loss of exergy.
However, in non-steady state the exergy balance of an open system reads as, cf.
Fig. 3.13:
dE x
= ṁex,S,1 + Ė x,Q + Pt −  Ė x,V − ṁex,S,2 . (16.86)
dt
The term on the left-hand side of the equation takes into account the change in
exergy in the system over time, which is influenced by the fluxes across the system
boundaries and the loss internally.
Overview open system
Figure 16.18 gives an overview of how open systems are treated in steady-state oper-
ation. It summarises what has been investigated so far:
• Law of mass conservation
ṁ 1 = ṁ 2 = ṁ (16.87)

Mass flux in is equal to mass flux out.

• First law of thermodynamics, see Fig. 16.18a

dE


= 0 = Q̇ + Pt + ṁ · h 1 + ea,1 − ṁ · h 2 + ea,2 (16.88)
dt
respectively


Q̇ + Pt + ṁ · h 1 + ea,1 = ṁ · h 2 + ea,2 (16.89)

Energy flux in is equal to energy flux out.

Partial energy equation


⎡ ⎤
2
Pt = ṁ · ⎣ v d p + ψ12 + ea,12 ⎦ (16.90)
1
404 16 Exergy

Fig. 16.18 Overview open system in steady state: balance of energy (a), entropy (b) and exergy
(c)

• Second law of thermodynamics, see Fig. 16.18b

dS Q̇
= 0 = ṁ · (s1 − s2 ) + + Ṡi (16.91)
dt Tm

Entropy flux in is equal to entropy flux out.


ṁs1 + + Ṡi = ṁs2 (16.92)
Tm

= Ṡa

• Balance of exergy, see Fig. 16.18c

dE x
= 0 = ṁex,S,1 + Ė x,Q + Pt −  Ė x,V − ṁex,S,2 (16.93)
dt
respectively
ṁex,S,1 + Ė x,Q + Pt =  Ė x,V + ṁex,S,2 (16.94)
16.4 Loss of Exerg 405

Fig. 16.19 Balance of


exergy: a clockwise cycle, b
counterclockwise cycle

Exergy flux in is equal to exergy flux out.

with
 Ė x,V = Tenv Ṡi (16.95)

16.4.3 Thermodynamic Cycles

In this section, the exergy loss in thermodynamic cycles is derived. Then an overview
of the underlying concepts is given. Moreover, a distinction is made between clock-
wise and counterclockwise cycles. Figure 16.19 shows the exergy fluxes for both
cycles. Since the cycles are operated in steady state,30 the exergy fluxes in must be
balanced by the fluxes out,31 i.e.
• Clockwise cycle, see Fig. 16.19a

Ė x,Q = P + Ė x,Q 0 +  Ė x,V . (16.96)

Hence, the loss of exergy is

 Ė x,V = Ė x,Q − Ė x,Q 0 − P. (16.97)

Substitution of the exergies of heat, cf. Eq. 16.9, results in


   
Tenv Tenv
 Ė x,V = Q̇ 1 − − Q̇ 0 1 − − P. (16.98)
T T0

30 The state of a system remains constant in steady state operation. Further assuming a constant
environment leads to a constant amount of exergy inside the cycle, i.e. a constant energetical distance
to the environment.
31 Note that  Ė
x,V is a sink.
406 16 Exergy

Rearranging leads to

Tenv Tenv
 Ė x,V = Q̇ − Q̇ 0 − P − Q̇ + Q̇ 0 . (16.99)
T T0

The first law of thermodynamics32 obeys

Q̇ = Q̇ 0 + P. (16.100)

Thus, Eq. 16.99 simplifies to


 
Q̇ 0 Q̇
 Ė x,V = Tenv − . (16.101)
T0 T

The second law of thermodynamics33 reads as

Q̇ Q̇ 0
+ Ṡi = . (16.102)
T T0

Hence, it finally is
 Ė x,V = Tenv Ṡi (16.103)

• Counterclockwise cycle, see Fig. 16.19b

Ė x,Q 0 + P = Ė x,Q +  Ė x,V . (16.104)

Hence, the loss of exergy is

 Ė x,V = Ė x,Q 0 − Ė x,Q + P. (16.105)

Substitution of the exergies of heat, see Eq. 16.9, results in


   
Tenv Tenv
 Ė x,V = Q̇ 0 1− − Q̇ 1 − + P. (16.106)
T0 T

Rearranging leads to

Tenv Tenv
 Ė x,V = Q̇ 0 − Q̇ + P − Q̇ 0 + Q̇ . (16.107)
T0 T

32 Energy flux in is balanced by energy flux out.


33 Entropy flux in is balanced by entropy flux out.
16.4 Loss of Exerg 407

The first law of thermodynamics obeys

Q̇ = Q̇ 0 + P. (16.108)

Thus, Eq. 16.107 simplifies to


 
Q̇ Q̇ 0
 Ė x,V = Tenv − . (16.109)
T T0

The second law of thermodynamics reads as

Q̇ 0 Q̇
+ Ṡi = . (16.110)
T0 T

Hence, it finally is
 Ė x,V = Tenv Ṡi (16.111)

Overview thermodynamic cycles


Figure 16.20 gives an overview of how clockwise cycles are treated in steady-state
operation. It summarises what has been investigated so far:
• First law of thermodynamics, cf. Fig. 16.20a

dU
= 0 = Q̇ − Q̇ 0 − P (16.112)
dt

Energy flux in is equal to energy flux out.

Fig. 16.20 Overview clockwise cycle: balance of energy (a), entropy (b) and exergy (c)
408 16 Exergy

Q̇ = Q̇ 0 + P (16.113)

• Second law of thermodynamics, see Fig. 16.20b

dS Q̇ Q̇ 0
=0= − + Ṡi (16.114)
dt T T0

Entropy flux in is equal to entropy flux out.

Q̇ Q̇ 0
+ Ṡi = (16.115)
T T0

• Balance of exergy, see Fig. 16.20c

Exergy flux in is equal to exergy flux out.

Ė x,Q = Ė x,Q 0 + P +  Ė x,V (16.116)

respectively
P = Ė x,Q − Ė x,Q 0 −  Ė x,V (16.117)

with
 Ė x,V = Tenv Ṡi (16.118)

• The exergetic efficiency is defined by the ratio of released power34 to the supplied
exergy flux. Thus, according to Fig. 16.20c it follows

released work P
ηex = = . (16.119)
supplied exergy Ė x,Q

Applying the balance of exergy, see Eq. 16.117, results in

P Ė x,Q − Ė x,Q 0 −  Ė x,V


ηex = = . (16.120)
Ė x,Q Ė x,Q

Rearranging leads to
Ė x,Q 0  Ė x,V
ηex = 1 − − . (16.121)
Ė x,Q Ė x,Q

34 Note that this is pure exergy.


16.4 Loss of Exerg 409


– The term Ėx,Q0 disappears in case the released heat flux Q̇ 0 does not contain any
x,Q
exergy, i.e. when T0 = Tenv
– The term ĖĖx,V disappears in case the machine works reversibly, i.e.  Ė x,V = 0
x,Q

In a theoretical limiting case, an exergetic efficiency of 100% is possible. Anyhow,


the thermal efficiency is defined as

P
ηth = . (16.122)

Expanding this equation leads to


P Ė x,Q
ηth = . (16.123)
Ė x,Q Q̇
  
=ηex

Substitution of the exergy of heat Ė x,Q results in


 
Tenv
ηth = ηex 1 − = ηex ηC (16.124)
T

The exergetic efficiency ηex evaluates the quality of the machine.35 In the best,
theoretical case, it can achieve 100% efficiency. This requires a reversible machine
without entropy generation and heat release at ambient temperature. However, the
Carnot efficiency ηC characterises the quality of the heat that drives the machine.
The higher the temperature at which the heat is available, the more exergy the heat
contains and the more efficient the energy conversion.

Figure 16.21 gives an overview of how counterclockwise cycles are treated in steady-
state operation. It summarises what has been investigated so far:

• First law of thermodynamics, see Fig. 16.21a

dU
= 0 = Q̇ 0 + P − Q̇ (16.125)
dt

Energy flux in is equal to energy flux out.

Q̇ 0 + P = Q̇ (16.126)

35 This efficiency is therefore relevant for the evaluation of the engineering performance.
410 16 Exergy

Fig. 16.21 Overview counterclockwise cycle: balance of energy (a), entropy (b) and exergy (c)

• Second law of thermodynamics, see Fig. 16.21b

dS Q̇ 0 Q̇
=0= + Ṡi − (16.127)
dt T0 T

Entropy flux in is equal to entropy flux out.

Q̇ 0 Q̇
+ Ṡi = (16.128)
T0 T

• Balance of exergy, see Fig. 16.21c

Exergy flux in is equal to exergy flux out.

Ė x,Q 0 + P = Ė x,Q +  Ė x,V (16.129)

respectively
P = Ė x,Q − Ė x,Q 0 +  Ė x,V (16.130)

with
 Ė x,V = Tenv Ṡi (16.131)

• The exergetic efficiency is:

Exergetic benefit
ηex = (16.132)
Exergetic effort
16.4 Loss of Exerg 411

– In case of a heat pump


Ė x,Q
ηex,HP = (16.133)
P
Applying the balance of exergy, see Eq. 16.130, results in

Ė x,Q
ηex,HP = (16.134)
Ė x,Q − Ė x,Q 0 +  Ė x,V

– In case of a cooling machine

Ė x,Q 0
ηex,CM = (16.135)
P
Applying the balance of exergy, cf. Eq. 16.130, results in

Ė x,Q 0
ηex,CM = (16.136)
Ė x,Q − Ė x,Q 0 +  Ė x,V

• The coefficient of performance is:

Benefit
ε= . (16.137)
Effort
– In case of a heat pump it follows


εHP = . (16.138)
P
The exergy of heat is  
Tenv
Ė x,Q = Q̇ 1 − . (16.139)
T

Hence, it follows

T Ė x,Q T
εHP = · = · ηex,HP . (16.140)
T − Tenv P T − Tenv

– In case of a cooling machine it follows

Q̇ 0
εCM = . (16.141)
P
The exergy of heat is  
Tenv
Ė x,Q 0 = Q̇ 0 1 − . (16.142)
T0
412 16 Exergy

Fig. 16.22 Illustration of energy fluxes

Hence, it follows

T0 Ė x,Q 0 T0
εCM = · = · ηex,CM . (16.143)
T0 − Tenv P T0 − Tenv

16.5 Sankey-Diagram

Changes of state in thermodynamic cycles or systems can be visualised with so-


called Sankey diagrams36 which represent the energy transformation in a flow chart.
Furthermore, energy can also be represented as exergy and anergy. The basic idea
is shown in Fig. 16.22. Exergy is the part of energy that can be converted into any
other form of energy, especially into mechanical or electrical work. In contrast to
exergy, anergy is the part that cannot be converted further and is therefore technically
useless. This is because it is already in equilibrium with the environment, so there is
no driving potential for conversion. Both exergy and anergy form the energy, which is
a conservation variable. It has been shown that any irreversibility in a process leads to
a loss of exergy and, to the same extent, an increase in anergy. In a reversible system,
however, no loss of exergy occurs. This is illustrated in Fig. 16.22. Energy enters a
system and with it both exergy and anergy. Due to the first law of thermodynamics,
the energy output in steady-state operation is equal to the energy input. However, if
irreversibilities occur within the process, exergy is lost and converted into anergy.

16.5.1 Open System

An example of an open system in steady state operation is shown in Fig. 16.23. A fluid
flow enters the system at the inlet (1). The entire flux of energy consists of the exergy

36Named after Matthew Henry Phineas Riall Sankey, November 1853 in Nenagh, Ireland, 3
October 1925.
16.5 Sankey-Diagram 413

Fig. 16.23 Sankey-diagram of an open system

of the flow ṁex,S,1 , see Sect. 16.2, and anergy of the flow ṁbx,S,1 . Depending on the
temperature at which the heat flux Q̇ is supplied to the system,37 cf. Sect. 16.1, an
exergy flux Ė x,Q as well as a flux of anergy Ḃx,Q enrich the system, see also Fig. 16.7.
At this stage, the total flux of energy is the sum of the flux of energy of the fluid flow
and the heat flux. In case technical power Pt is additionally supplied, the entire flux
of energy rises. This also applies to the exergy flux, because technical power is pure
exergy. However, if there are imperfections internally, such as dissipation or other
inhomogeneities, a loss of exergy flux appears. Thus, the exergy flux is reduced,
while the flux of anergy rises by that amount. Finally, the flux of energy leaving at
the outlet (2) is exactly the same as the initial flux of energy at the inlet plus the
fluxes of energy, that have been supplied across the system boundary. However, this
flux of energy is composed of exergy of the flow Ė x,S,2 and anergy of the flow Ḃx,S,2 .

16.5.2 Heat Transfer

Figure 16.24 shows an example of steady-state heat transfer through a conductive


wall. This problem has been discussed several times, cf. Sect. 14.2, as it is an excellent
example of how entropy can be generated even though there is no dissipation, i.e.
internal friction. The cause of a heat flux is a temperature difference, for example
T1 > T2 , as shown in the figure. From a thermodynamic point of view, the wall is an
imperfection as it inhibits to reach a thermodynamic equilibrium between left and
right side, i.e. a homogeneous temperature distribution. To achieve this equilibrium,38

37 The system boundary only includes the fluid, i.e. no exergy loss due to heat transfer via the heat
exchanger housing is taken into account.
38 However, as long as T and T are fixed, i.e. by heating/cooling, an equilibrium can not be reached.
1 2
414 16 Exergy

Fig. 16.24 Heat transfer (wall)

a heat flux occurs that follows the temperature gradient according to the second law
of thermodynamics. The heat flux at any horizontal position39 must be the same. If
this were not the case, the internal energy would vary spatially and temporally, i.e.
it would not be a steady-state problem. Since the heat flux at the right edge is equal
to the heat flux leaving the system at the left edge, the incoming entropy flux due to
heat is smaller than the outgoing entropy flux due to heat. Consequently, a flux of
entropy generation must take place within the wall, which is the cause of the whole
process being irreversible: Once the heat flux has left the system, there is no way
back by itself.40 The rate of entropy generation has been derived already and follows

T1 − T2
Ṡi = Q̇ . (16.144)
T1 T2

This correlation shows, that the heat transfer mechanism is only reversible, if a heat
flux is theoretically transferred at T → 0.
An exergy flux of heat passes into the system on the right hand side. Due to the
high temperature of the heat flux, the exergy flux on the right hand side is greater than
the exergy flux, that leaves on the left hand side at a lower temperature. Obviously,

39 In case of one-dimensional heat transfer.


40 Bringing back the heat against the gradient requires a heat pump, i.e. a counterclockwise cycle,
that consumes technical power.
16.5 Sankey-Diagram 415

the working capability of the heat flux has decreased while passing the wall, i.e. there
is a loss of exergy flux internally, which obeys

T1 − T2
 Ė x,V = Tenv Q̇ = Tenv Ṡi . (16.145)
T1 T2

Problem 16.2 Through the wall of a house with an internal temperature of ϑ =


22 ◦ C, a heat flux of Q̇ = 20 kW is released to the environment with a temperature
of ϑenv = −5 ◦ C. The process shall be investigated in steady state.
(a) What is the entropy flux Ṡi , that is generated in this process?
(b) What loss of exergy flux  Ė x,V occurs?
(c) What would be the minimum power consumption of a heat pump required to
heat the house under these conditions?

Solution

Figure 16.25 shows the sketches that are required for parts (a) and (b).

(a) This part has already been solved in Problem 14.2. A balance of entropy in steady
state leads to
Q̇ Q̇
+ Ṡi = . (16.146)
T1 Tenv

The flux of entropy generation yields


 
1 1 W
Ṡi = Q̇ − = 6.82 > 0. (16.147)
Tenv T1 K

The process is therefore irreversible, i.e. once the heat flux has left the building,
it does not return due to the temperature gradient.
(b) There are two possibilities to calculate the loss of exergy flux  Ė x,V :
• Based on the generation of entropy:

Fig. 16.25 Sketch for the solution to Problem 16.2


416 16 Exergy

 Ė x,V = Tenv Ṡi = 1.8296 kW (16.148)

• Based on a balance of exergy:


According to Fig. 16.25 the exergy balance in steady state41 obeys

Ė x,Q@T1 =  Ė x,V + Ė x,Q@Tenv . (16.149)

The loss of exergy flux follows accordingly

 Ė x,V = Ė x,Q@T1 − Ė x,Q@Tenv (16.150)

with the exergy fluxes of heat


 
Tenv
Ė x,Q@T1 = Q̇ 1 − = 1.8296 kW (16.151)
T1

and  
Tenv
Ė x,Q@Tenv = Q̇ 1 − = 0. (16.152)
Tenv

Thus, the entire exergy flux of state (1) is lost, i.e.

 Ė x,V = Ė x,Q@T1 = 1.8296 kW. (16.153)

(c) Again, there are several alternatives according to Fig. 16.26.


• Based on first and second law of thermodynamics:

Q̇ env + P = Q̇. (16.154)

Hence, the minimum power (in case the heat pump is operated reversibly)
follows
P = Q̇ − Q̇ env . (16.155)

In order to find Q̇ env the second law of thermodynamics is applied, i.e.

Q̇ env Q̇
+ Ṡi = . (16.156)
Tenv  T1
=0

Thus, the heat flux taken from the environment results in


Q̇ env = Tenv = 18.17 kW. (16.157)
T1

41 Exergy in is balanced by exergy out. Note that the exergy loss reduces the exergy of a system. It
therefore counts as outgoing exergy.
16.5 Sankey-Diagram 417

Fig. 16.26 Sketch for the solution to Problem 16.2

Consequently, the required power is

P = Q̇ − Q̇ env = 1.8296 kW. (16.158)

• Based on a balance of exergy according to system boundary A42 :

Ė x,Q env @Tenv + P = Ė x,Q@T1 +  Ė x,V (16.159)


  
=0

Since Q̇ env occurs at Tenv it is

Ė x,Q env @Tenv = 0. (16.160)

Hence, the required power P reads as


 
Tenv
P = Ė x,Q@T1 = Q̇ 1 − = 1.8296 kW. (16.161)
T1

• Based on a balance of exergy according to system boundary B:

Ė x,Q env @Tenv + P = Ė x,Q@Tenv +  Ė x,V + Ė x,V,b) (16.162)


  
=0

In this equation  Ė x,V,b) is the loss of exergy flux within the wall according
to part (b), see Figs. 16.25 and 16.26. Since

42 The loss of exergy flux in part (b) has nothing to do with the loss of exergy flux in part (b), since
the system boundary is different.
418 16 Exergy

Ė x,Q@Tenv = 0 (16.163)

respectively
Ė x,Q env @Tenv = 0 (16.164)

the power finally results in


P =  Ė x,V,b) . (16.165)

Problem 16.3 Air is compressed in an adiabatic compressor in a steady state process


from state (1) ( p1 = 1.05 bar, ϑ1 = ϑenv = 15 ◦ C) to a pressure of p2 = 6.25 bar
in state (2). Air can be treated as an ideal gas with an individual gas constant of
R = 0.287 kgkJK and a constant specific heat capacity of c p = 1.004 kgkJK . A specific
technical work of wt = 230 kgkJ
is required.
(a) What is the specific loss of exergy ex,V,12 ?
(b) Sketch the change of state in a T, s-diagram and mark the specific dissipated
energy by a vertical hatching and the specific exergy loss ex,V,12 by a horizontal
hatching.
(c) Calculate the specific dissipated energy.

Solution

Figure 16.27 shows the required balances.


(a) The specific loss of exergy ex,V,12 follows

ex,V = Tenv si,12 . (16.166)

In order to calculate the specific generation of entropy, i.e. si,12 , a balance of


entropy is applied:
s1 + si,12 + sa,12 = s2 . (16.167)

=0

Fig. 16.27 Sketch for the solution to Problem 16.3


16.5 Sankey-Diagram 419

The specific generation of entropy obeys

T2 p2
si,12 = s2 − s1 = c p ln − R ln . (16.168)
T1 p1

T2 is unknown, so that the first law of thermodynamics is applied, i.e.

q12 +wt12 = h 2 − h 1 = c p (T2 − T1 ) . (16.169)



=0

This leads to the outlet temperature T2 , i.e.


wt12
T2 = T1 + = 517.23 K. (16.170)
cp

For the specific generation of entropy it follows accordingly

T2 p2 kJ
si,12 = c p ln − R ln = 0.0754 . (16.171)
T1 p1 kg K

Thus, the loss of exergy finally is

kJ
ex,V = Tenv si,12 = 21.726 . (16.172)
kg

Alternatively, a balance of exergy can be performed, i.e.

ex,S,1 + ex,Q +wt12 = ex,S,2 + ex,V . (16.173)



=0

Hence, the specific loss of exergy follows

ex,V = ex,S,1 − ex,S,2 + wt12 . (16.174)

The specific exergies of the flow can be substituted by

1
ex,S,1 = h 1 − h env + c12 + gz 1 + Tenv (senv − s1 ) (16.175)
2
respectively

1
ex,S,2 = h 2 − h env + c22 + gz 2 + Tenv (senv − s2 ) . (16.176)
2
420 16 Exergy

Finally the specific loss of exergy is

1
2
ex,V = h 1 − h 2 + c1 − c22 + g (z 1 − z 2 ) + Tenv (s2 − s1 ) + (h 2 − h 1 ) .
2   
=wt12
(16.177)
Consequently, the loss of exergy finally results in

T2 p2 kJ
ex,V = Tenv (s2 − s1 ) = Tenv c p ln − R ln = 21.726 . (16.178)
T1 p1 kg

(b) Figure 16.28 shows the solution.


(c) The specific dissipated energy can be calculated with the partial energy equation,
i.e.
2
kJ
wt12 = 230 = v d p + ψ12 + ea12 . (16.179)
kg   
1 =0

Hence, the specific dissipation results in

2
ψ12 = wt12 − v dp . (16.180)
1
  
=y12

The change of state is polytropic.43 Thus, it can be described by a polytropic


exponent n. The pressure work y12 for a polytropic change of state is

RT1 T2
y12 = n −1 . (16.181)
n − 1 T1

The specific dissipation results in



RT1 T2
ψ12 = wt12 − n −1 . (16.182)
n − 1 T1

In order to find the exponent n the polytropic change of state is investigated, i.e.

p1 v1n = p2 v2n . (16.183)

Thus, the exponent n obeys


ln pp21
n= . (16.184)
ln vv21

43 Note that every change of state is polytropic: In this case it is an adiabatic and a frictional change.
16.5 Sankey-Diagram 421

Fig. 16.28 Sketch for the


solution to Problem 16.3

The specific volumes v1 and v2 follow the thermal equation of state, i.e.

RT1 m3
v1 = = 0.7876 (16.185)
p1 kg

respectively
RT2 m3
v2 = = 0.2375 . (16.186)
p2 kg

Hence, the exponent n is


ln pp21
n= = 1.488. (16.187)
ln vv21

Finally, the specific dissipation results in



RT1 T2 kJ
ψ12 = wt12 − n − 1 = 29.5278 . (16.188)
n − 1 T1 kg
Chapter 17
Components and Thermodynamic Cycles

Although many technical components required for the operation of energy conver-
sion processes have been discussed previously,1 the focus in this chapter is on thermal
turbomachinery as well as on heat exchangers. However, in this chapter the technical
components are addressed in steady state operation. In order to quantify the effi-
ciency of turbine and compressor, the so-called isentropic efficiency is defined. In
addition, relevant thermodynamic cycles are presented and discussed. Cyclic pro-
cesses have also been treated in the previous chapters, but mostly in a black-box
notation.2 A distinction is made between clockwise cycles, i.e. thermal engines, and
counterclockwise cycles, i.e. refrigerators or heat pumps. This chapter concludes
Part I of this book.

17.1 Components

17.1.1 Turbine

A turbine is a thermal turbomachine that converts the enthalpy of a fluid into rotational
energy, i.e. mechanical work is released, see Fig. 17.1. Thereby the pressure of the
fluid decreases from p1 to p2 < p1 .
• The first law of thermodynamics for a turbine in steady state reads as3 :

1 Such as adiabatic throttle, compressor and turbine.


2 With the exception of the Carnot cycle, i.e. the technical benchmark, whose underlying changes
of state, i.e. isentropic, isothermal, isothermal, have also been introduced, see Sect. 15.3.
3 It applies that the energy flux in and the energy flux out are in equilibrium.

© Springer Nature Switzerland AG 2022 423


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_17
424 17 Components and Thermodynamic Cycles

Fig. 17.1 Turbine: a First law, b Second law, c Exergy

   
1 2 1 2
Q̇ 12 + Pt12 + ṁ h 1 + c1 + gz 1 = ṁ h 2 + c2 + gz 2 . (17.1)
2 2

In specific notation and neglecting the potential energy, which is assumed to be


comparatively small in technical applications, it results that

1 2 
q12 + wt12 = h 2 − h 1 + c2 − c12 . (17.2)
2
With the so-called specific total enthalpy

c2
h+ = h + (17.3)
2
the first law of thermodynamics finally obeys

q12 + wt12 = h + +
2 − h1 (17.4)

• The partial energy equation reads as

2
1 2 
wt,12 = v d p + ψ12 + c2 − c12 + g (z 2 − z 1 ) . (17.5)
2
1

Ignoring the potential energy simplifies the partial energy equation, i.e.
17.1 Components 425

2
1 2 
wt,12 = v d p + ψ12 + c2 − c12 (17.6)
2
1

• The second law of thermodynamics in steady state4 is, see Fig. 17.1b,

ṁs1 + Ṡi + Ṡa = ṁs2 . (17.7)

In specific notation, by dividing with the mass flow rate, it is

s2 − s1 = si12 + sa12 (17.8)

In differential notation it obeys

δψ δq
ds = δsi + δsa = + . (17.9)
T

T
≥0

• The balance of exergy in steady state5 reads as, cf. Fig. 17.1c,

ṁex,S,1 + Ė x,Q + Pt12 = ṁex,S,2 +  Ė x,V (17.10)

with
 Ė x,V = Tenv Ṡi . (17.11)

Adiabatic turbine

For an adiabatic turbine with neglected potential and kinetic energy, the first law of
thermodynamics is simplified as follows

wt12 = h 2 − h 1 < 0. (17.12)

In order for the turbine to release as much specific technical work as possible, the
difference between the specific enthalpies of (1) and (2) must be as large as possible.
However, the second law of thermodynamics under these premises follows

s2 − s1 = si12 ≥ 0. (17.13)

4 Entropy flux in is balanced by entropy flux out.


5 Exergy flux in is balanced by exergy flux out.
426 17 Components and Thermodynamic Cycles

Fig. 17.2 Adiabatic turbine:


Illustration in a T, s-diagram
(ideal and real gases)

From state (1) to state (2), the specific entropy can either remain constant or it must
increase. Figure 17.2 shows the change of state in a T, s-diagram. As the pressure
decreases from (1) to (2), state (2) is at a lower isobaric level than state (1). The
second law of thermodynamics has further shown, that since no heat is exchanged,
the specific entropy needs to rise in case dissipation occurs:

ψ12 → si12 → ex,V . (17.14)

Consequently, state (2) lies below and to the right of state (1). In best case, i.e. the
change of state runs free of dissipation, the change of state is reversible and adiabatic,
i.e. isentropic, so that the change of state runs vertically in a T, s-diagram. Under
isentropic conditions, i.e. in the best case, the state (2s) is reached. The T, s-diagram
shown additionally contains the isenthalps, i.e. curves of constant specific enthalpy.
For an ideal gas, the isenthalps are horizontal because the specific enthalpy is solely
a function of the temperature, i.e.

dh = c p dT. (17.15)

Figure 17.2, however, also contains the isenthalps for real fluids, which are dealt with
in Part II. According to Eq. 12.99, the correlation between temperature T , specific
enthalpy h and pressure p is more complex for real fluids than for ideal gases.
Nevertheless, the first law of thermodynamics has shown that the specific technical
work is a function of h = h 2 − h 1 . Figure 17.2 thus proves that the thermodynamic
process from (1) to (2) is not yet optimised, as h can be further increased. Obviously,
the maximum technical work in an adiabatic turbine within a pressure range p1 . . . p2
can be realised by an isentropic change of state (1) → (2s):
• (1) → (2): Adiabatic, technically realistic turbine, i.e. irreversible. The technical
power is not maximised.
17.1 Components 427

• (1) → (2s): Adiabatic, hypothetically best turbine, i.e. reversible.6 Maximum


technical power can be gained by this isentropic change of state.
Therefore, it is reasonable to compare these two adiabatic processes on the basis of
the so-called isentropic efficiency7 ηsT , i.e.

h h2 − h1
ηsT = = (17.16)
h s h 2s − h 1

In case the gas is ideal, the definition simplifies to

h T2 − T1
ηsT = = . (17.17)
h s T2s − T1

Thus, the power of the turbine can be calculated as follows if the turbine is adiabatic
and kinetic/potential energies are ignored:

Pt12 = ṁwt12 = ṁ h = ṁηsT h s < 0. (17.18)

Typical values for isentropic efficiencies are, see [5]:


• Steam turbine: 0.88 . . . 0.94
• Gas turbine: 0.90 . . . 0.95

17.1.2 Compressor

A compressor is a thermal turbomachinery that compresses a fluid from a pressure


p1 to a pressure p2 > p1 by supplying work, see Fig. 17.3. The following balances
can be conducted according to the turbine:
• The first law of thermodynamics for a compressor in steady state reads as8 :
   
1 1
Q̇ 12 + Pt12 + ṁ h 1 + c12 + gz 1 = ṁ h 2 + c22 + gz 2 . (17.19)
2 2

In specific notation and neglecting the potential energy, which is supposed to be


comparatively small in technical applications, it results that

1 2 
q12 + wt12 = h 2 − h 1 + c − c12 . (17.20)
2 2

6 So there is no dissipation or other imperfections that lead to the generation of entropy.


7 Such a definition ensures that the efficiency is always less than 1.
8 It applies that the energy flux in and the energy flux out are in equilibrium.
428 17 Components and Thermodynamic Cycles

Fig. 17.3 Compressor: a First law, b Second law, c Exergy

With the specific total enthalpy the first law of thermodynamics finally obeys

q12 + wt12 = h + +
2 − h1 (17.21)

• The partial energy equation reads as:

2
1 2 
wt,12 = v d p + ψ12 + c2 − c12 + g (z 2 − z 1 ) . (17.22)
2
1

Ignoring the potential energy simplifies the partial energy equation, i.e.

2
1 2 
wt,12 = v d p + ψ12 + c2 − c12 (17.23)
2
1

• The second law of thermodynamics in steady state9 is, see Fig. 17.3b,

ṁs1 + Ṡi + Ṡa = ṁs2 . (17.24)

In specific notation, by dividing with the mass flow rate, it is

s2 − s1 = si12 + sa12 (17.25)

9 Entropy flux in is balanced by entropy flux out.


17.1 Components 429

In differential notation it obeys

δψ δq
ds = δsi + δsa = + . (17.26)
T

T
≥0

• The balance of exergy in steady state10 reads as, see Fig. 17.3c,

ṁex,S,1 + Ė x,Q + Pt12 = ṁex,S,2 +  Ė x,V (17.27)

with
 Ė x,V = Tenv Ṡi . (17.28)

Adiabatic compressor

For an adiabatic compressor with neglected potential and kinetic energy, the first law
of thermodynamics is simplified to

wt12 = h 2 − h 1 > 0. (17.29)

In order to minimise the specific technical work required, the difference between the
specific enthalpies of (1) and (2) must therefore be as small as possible. However,
under these premises, the second law of thermodynamics reads as follows

s2 − s1 = si12 ≥ 0. (17.30)

From state (1) to state (2), the specific entropy can therefore either remain constant
or it must increase. Figure 17.4 shows the change of state in a T, s-diagram. As the
pressure increases from (1) to (2), state (2) is at a higher isobaric level than state (1).
The second law of thermodynamics has shown that since no heat is exchanged, the
specific entropy must increase when dissipation occurs:

ψ12 → si12 → ex,V . (17.31)

Consequently, state (2) lies above and to the right of state (1). In best case, i.e. the
change of state is free of dissipation, the change of state is reversible and adiabatic,
i.e. isentropic, so that the change of state runs vertically in a T, s-diagram. Under
isentropic conditions, i.e. in the best case, the state (2s) is reached. The T, s-diagram
shown additionally contains the isenthalps. Figure 17.4 not only shows the isenthalp
for an ideal gas, but it also contains the isenthalps for real fluids. However, the first
law of thermodynamics has shown that the specific technical work is a function of

10 Exergy flux in is balanced by exergy flux out.


430 17 Components and Thermodynamic Cycles

Fig. 17.4 Adiabatic


compressor: Illustration in a
T, s-diagram (ideal and real
gases)

h = h 2 − h 1 . Figure 17.4 thus proves that the thermodynamic process from (1) to
(2) is not yet optimised, as h can be further reduced. Obviously, the minimum
technical work in an adiabatic compressor within a pressure range p1 . . . p2 can be
realised by an isentropic change of state (1) → (2s):
• (1) → (2): Adiabatic, technically realistic compressor, i.e. irreversible. The tech-
nical power is not minimised.
• (1) → (2s): Adiabatic, hypothetically best compressor, i.e. reversible.11 Minimum
technical power can be achieved by this isentropic change of state.
Therefore, it is reasonable to compare these two adiabatic processes on the basis of
the so-called isentropic efficiency12 ηsC , i.e.

h s h 2s − h 1
ηsC = = (17.32)
h h2 − h1

In case the gas is ideal, the definition simplifies to

h s T2s − T1
ηsC = = . (17.33)
h T2 − T1

The power of the compressor can therefore be calculated as follows if the compressor
is adiabatic and kinetic/potential energies are ignored:

ṁ h s
Pt12 = ṁwt12 = ṁ h = > 0. (17.34)
ηsC

Typical values for isentropic efficiencies are in the range of 0.85 . . . 0.90, see [5].

11 So there is no dissipation or other imperfections that lead to the generation of entropy.


12 Such a definition ensures that the efficiency is always less than 1.
17.1 Components 431

Fig. 17.5 Thermal turbo-machines in a h, s-diagram

17.1.3 Thermal Turbomachines in a h, s-Diagram

Instead of a T, s-diagram, the previous considerations have shown that it is advan-


tageous to illustrate changes of state in a h, s-diagram. Hence, Fig. 17.5 summarises
the changes of state in an adiabatic turbine respectively in an adiabatic compressor. If
the change of state is adiabatic, the first law of thermodynamics in specific notation
is as follows

1 2   
wt12 = (h 2 − h 1 ) + c − c12 + g (z 2 − z 1 ) = h + +
2 − h 1 + g (z 2 − z 1 ) .
2 2
(17.35)
By applying the specific total enthalpy, see Eq. 17.3, it is

  < 0 adiabatic turbine
wt12 = h+
2 − h+
1 + g (z 2 − z 1 ) (17.36)
> 0 adiabatic compressor.

Under these conditions, the specific technical work wt12 can be easily visualised in
a h, s-diagram as shown in Fig. 17.5. If the kinetic respectively potential
 energy is 
also taken into account, the specific enthalpy h must be extended by 0.5c2 + gz ,
as done in Fig. 17.5. In general, the outer energies are often of secondary importance
compared to the caloric energy, so that they can be neglected in a steam power
plant, for example. As already mentioned, for an ideal gas a h, s-diagram can be
easily adapted from a T, s-diagram, since specific enthalpy h is proportional to
temperature T . Consequently, isobars and isochors lines have the same mathematical
shape as they have in a T, s-diagram. However, a h, s-diagram for non-ideal, i.e. real,
gases/fluids is introduced in Part II.
432 17 Components and Thermodynamic Cycles

17.1.4 Adiabatic Throttle

An adiabatic throttle is utilised to reduce the pressure of a fluid. Unlike a turbine,


which is also used to reduce pressure, no technical work is released. A throttle is
therefore a passive, i.e. work-isolated component. The working principle is illustrated
in Fig. 17.6. A pressure drop is simply caused by a reduction in the cross-section:
this triggers turbulence, i.e. dissipation. It has already been shown that dissipation
generates entropy. From a technical point of view, a throttle is a component built to
dissipate energy, while other technical applications increase their efficiency by avoid-
ing dissipation. An adiabatic throttle has already been calculated in Problem 12.5.
However, the following assumptions and premises are made:
(1) No technical work is exchanged,13 i.e. wt12 = 0
(2) The throttle is adiabatic, i.e. q12 = 0.
(3) The throttle is operated in steady state, i.e. ṁ 1 = ṁ 2 = ṁ.
(4) The change of potential energy is neglected, i.e. z 1 = z 2 .
(5) The change of kinetic energy is neglected, i.e. c12 = c22 .
The thermodynamic balances are as follows:
• First law of thermodynamics

1 2 
q12 + wt12 = h 2 − h 1 + c2 − c12 + g · (z 2 − z 1 ) (17.37)
2
With the above assumptions, the first law of thermodynamics is simplified to

h2 − h1 = 0 (17.38)

In other words, an adiabatic throttle is isenthalpic, i.e. dh = 0.


• Partial energy equation14

2
wt12 = v d p + ψ12 = 0 (17.39)
1

Rearranging leads to
2
v d p = −ψ12 < 0 (17.40)
1

13 There are no mechanical/electrical parts within the component, that exchange work with the
environment.
14 Outer energies are ignored.
17.1 Components 433

Fig. 17.6 Adiabatic throttle

It is therefore the dissipation that causes the pressure loss in an adiabatic throttle
under the assumptions made.

• Second law of thermodynamics

ṁs1 + Ṡa + Ṡi = ṁs2 (17.41)

According to the second law of thermodynamics, in steady state the entropy flux
into the system must be balanced by the entropy flux out of the system. Since the
throttle is to be adiabatic, the second law of thermodynamics obeys

ṁs1 + Ṡi = ṁs2 (17.42)

respectively
Ṡi = ṁ (s2 − s1 ) (17.43)

• Balance of exergy

ṁex,S,1 + Ė x,Q + Pt12 = ṁex,S,2 +  Ė x,V (17.44)




=0

In steady state, the exergy flux into the system must be balanced by the released
flux. Note that the loss of exergy flux due to the flux of entropy generation is a
sink from a thermodynamic point of view, while the flux of entropy generation is
a source. However, since the throttle is adiabatic, there is no exergy flux of heat,
so that the balance simplifies to

ṁex,S,1 = ṁex,S,2 +  Ė x,V (17.45)

As already mentioned, dissipation is the cause of the pressure loss in an adiabatic


throttle. It is now derived how much energy is dissipated. However, there are two
alternative approaches, both of which are presented:
434 17 Components and Thermodynamic Cycles

• Approach following the partial energy equation

2
wt12 = v d p + ψ12 = 0 (17.46)
1

The partial energy equation can be solved for the specific dissipation, i.e.

2
ψ12 = − v d p. (17.47)
1

To solve the integral, a correlation between pressure p and specific volume v is


required. Since the first law of thermodynamics has shown that the change of state
is isenthalpic, it follows for an ideal gas that

h 2 = h 1 ⇒ T2 = T1 . (17.48)

Thus, applying the thermal equation of state reads as

pv = RT = const. (17.49)

respectively
p1 v1
pv = p1 v1 → v = . (17.50)
p

The substitution in the partial energy equation leads to

2 2 2
1 1
ψ12 = − v d p = − p1 v1 d p = −RT d p. (17.51)
p p
1 1 1

Hence, it finally is
p2
ψ12 = −RT ln (17.52)
p1

• Approach following the second law of thermodynamics

ṁ (s1 − s2 ) + Ṡa12 + Ṡi12 = 0 (17.53)

Since the throttle is adiabatic, the following results in specific notation

s2 − s1 = si12 . (17.54)
17.1 Components 435

The application of the caloric equation of state yields

T2 p2
s2 − s1 = c p · ln − R · ln . (17.55)
T1 p1

Since for an ideal gas an isenthalpic change of state is also isothermal, the second
law of thermodynamics can be simplified, i.e.
p2
s2 − s1 = −R · ln = si12 . (17.56)
p1

The specific generation of entropy obeys

2
δψ
si12 = . (17.57)
T
1

Hence, it follows
2
p2 δψ
− R · ln = . (17.58)
p1 T
1

Since the temperature is constant, the integral can be simplified, i.e.

2
p2 1 ψ12
− R · ln = δψ = . (17.59)
p1 T T
1

By rearranging, one finally obtains the specific dissipation, i.e.

p2
ψ12 = −RT ln (17.60)
p1

17.1.5 Heat Exchanger

A heat exchanger is a caloric apparatus, that is passed by two or even more fluids
with different temperatures, i.e. in thermal disequilibrium.15 In order to achieve a
thermal equilibrium, heat is exchanged between the fluids. However, heat exchangers
have already been treated in several problems before. There are many technical
applications, e.g. heater, boiler, oil- and water coolers, that require heat exchangers. A
distinction can be made according to the type of heat exchanger: plate heat exchanger,
shell and tube heat exchanger, adiabatic wheel heat exchanger and several others

15 Such a disequilibrium causes a transient balancing process.


436 17 Components and Thermodynamic Cycles

Fig. 17.7 Symbols for heat exchangers

Fig. 17.8 Counterflow heat


exchanger

more. The focus within this book is on heat exchangers that do not allow a mixing
of the fluids. Anyhow, heat exchangers can be operated in parallel or counterflow.
Possible technical symbols for heat exchangers are shown in Fig. 17.7.

Energy balancing heat exhangers

Let us take a closer look at the heat exchanger shown in Fig. 17.8, which consists of
two flows in counterflow operation. In this case, the first law of thermodynamics in
steady state, i.e. the energy flux in is balanced by the energy flux out, is as follows

Q̇ iso + Pt + ṁ 1 h 1,E + c1,E
2
/2 + g · z 1,E + ṁ 2 h 2,E + c2,E
2
/2 + g · z 2,E

= ṁ 1 h 1,A + c1,A
2
/2 + g · z 1,A + ṁ 2 h 2,A + c2,A
2
/2 + g · z 2,A .
(17.61)
Since the heat exchanger is a passive component, i.e. no work is transferred, the
first law of thermodynamics can be simplified. Let us further assume that the heat
exchanger does not exchange any heat with the environment, i.e. Q̇ iso = 0, and that
the potential energy can be neglected. Thus, the first law of thermodynamics obeys
+
0 = ṁ 1 h + + +
1,A − h 1,E + ṁ 2 h 2,A − h 2,E . (17.62)

Note that h + is the total specific enthalpy according to Eq. 17.3, which includes
both the specific caloric enthalpy and the specific kinetic energy. The reformulation
leads to +
ṁ 1 h + + +
1,A − h 1,E = ṁ 2 h 2,E − h 2,A (17.63)
17.1 Components 437

Fig. 17.9 Counterflow heat


exchanger

respectivley, if kinetic energy is ignored,



ṁ 1 h 1,A − h 1,E = ṁ 2 h 2,E − h 2,A (17.64)

Obviously, the balance according to the system boundary in Fig. 17.8 does not
contain any information about the heat transferred between the fluids. Alternatively,
the first law of thermodynamics is now applied for the system boundaries illustrated
in Fig. 17.9. As a result, the heat Q̇ 12 transferred between the two fluids is now
included in the balance. The first law of thermodynamics in steady state for Fluid 1,
see Fig. 17.9, reads under the same assumptions as before

Q̇ 12 + ṁ 1 h 1,E = ṁ 1 h 1,A . (17.65)

The application of the caloric equation of state for ideal gases results in

Q̇ 12 = ṁ 1 c p,1 T1,A − T1,E . (17.66)

The first law of thermodynamics in steady state for Fluid 2, see Fig. 17.9 is

Q̇ 12 + ṁ 2 h 2,A = ṁ 2 h 2,E . (17.67)

The application of the caloric equation of state for ideal gases results in

Q̇ 12 = ṁ 2 c p,2 T2,E − T2,A . (17.68)

However, the combination of Eqs. 17.65 and 17.67 leads to Eq. 17.64.

Counterflow versus parallel flow

As of now, heat exchangers have been depicted as counterflow systems. For the
application of the first law of thermodynamics, the direction of flow is irrelevant as
438 17 Components and Thermodynamic Cycles

Fig. 17.10 Temperature profiles in a heat exchanger: a Parallel flow b Counterflow

Fig. 17.11 Entropy balance (heat exchanger is supposed to be adiabatic to the environment)

long as the rule is kept that the energy flow in is balanced by the energy flow out.
Nevertheless, Fig. 17.10 shows a comparison between parallel and counterflow. In
general, a counterflow heat exchanger operated at the same inlet temperatures of the
fluids as a parallel flow heat exchanger is more efficient than a parallel flow heat
exchanger. This is due to the larger average temperature difference compared to a
parallel flow.

Entropy balancing heat exchangers

The entropy balance for heat transfer has already been investigated in Chap. 14 for a
conductive wall. It has been shown that any heat transfer at a temperature difference
greater than zero is accompanied by the generation of entropy. In this section, the
entropy balance for an entire heat exchanger, see Fig. 17.11, is applied. If only system
A is examined, the entropy generation in A is exclusively due to dissipation: If kinetic
and potential energies are ignored, the partial energy equation simplifies to
17.1 Components 439

2
wt = 0 = v d p + ψ12 . (17.69)
1

The pressure drop d p < 0 is thus an indicator of the entropy generated, i.e.

2
ψ12 = − v d p. (17.70)
1

A steady state balance of entropy, i.e. flux of entropy into the system is balanced by
the flux of entropy out of the system, for system A yields


ṁ 1 s1 + Ṡi,A = ṁ 1 s2 + . (17.71)
Tm,A

If there is no pressure drop, under the assumptions given, there is also no flux of
entropy generation Ṡi,A .16

Nevertheless, the thermodynamic mean temperature for the system is, see
Sect. 13.7.3,
q12 + ψ12
Tm,A = . (17.72)
s2 − s1

Applying the first law of thermodynamics

q12 = h 1 − h 2 (17.73)

leads to
h 1 − h 2 + ψ12
Tm,A = . (17.74)
s2 − s1

Thus, for steady state it is

dSA Q̇
= 0 = Ṡi,A − − ṁ 1 (s2 − s1 ) (17.75)
dt Tm,A

According to Eq. 17.75, the entropy within system A is temporally constant,17 i.e.
the flux in is balanced by the flux out. Now, let us focus on system B according to
Fig. 17.11. The balance of entropy for system B obeys

16 Kinetic and potential energies are not considered.


17 Thus, it is dSdtA = 0.
440 17 Components and Thermodynamic Cycles

Fig. 17.12 Entropy balance (system C)


ṁ 2 s3 + Ṡi,B + = ṁ 2 s4 . (17.76)
Tm,B

The thermodynamic mean temperature for system B follows

q34 + ψ34
Tm,B = . (17.77)
s4 − s3

Applying the first law of thermodynamics

q34 = h 4 − h 3 (17.78)

leads to
h 4 − h 3 + ψ34
Tm,B = . (17.79)
s4 − s3

Thus, for steady state it is

dSB Q̇
= 0 = Ṡi,B + − ṁ 2 (s4 − s3 ) (17.80)
dt Tm,B

However, the overall system additionally consists of a third system C, i.e. a wall
separating the two fluids, cf. Fig. 17.12. The balance of entropy for system C results
in

Q̇ Q̇
Ṡi,C + = . (17.81)
Tm,A Tm,B

In steady state operation the heat flux into system C must be equal to the heat flux
out of system C. However, the thermodynamic mean temperatures of systems A and
B can be measured directly at the system boundary of the wall,18 see Fig. 17.12b.
Consequently, the entropy flux transferred by the heat flux can be easily calculated,

18 Systems A and B are homogenised by their thermodynamic mean temperatures.


17.1 Components 441

Fig. 17.13 Overall entropy balance

see Eq. 17.81. The generated entropy flux in the wall, i.e. in system C, obeys19 :

Q̇ Q̇
Ṡi,C = − . (17.82)
Tm,B Tm,A

Thus, in an alternative notation it follows

dSC Q̇ Q̇
= 0 = Ṡi,C + − (17.83)
dt Tm,A Tm,B

The overall system is the summation of the three subsystems A, B and C, see
Fig. 17.13. Again, in steady state, the entropy in the overall system must be con-
stant in time, so that the summation of systems A, B and C yields
dStotal dSA dSB dSC
=0= + + . (17.84)
dt dt dt dt
The substitution of the previously derived equations results in

0 = Ṡi,A + Ṡi,B + Ṡi,C −ṁ 1 (s2 − s1 ) − ṁ 1 (s4 − s3 ) . (17.85)


 

= Ṡi,total

Thus, it finally is
Ṡi,total = ṁ 1 (s2 − s1 ) + ṁ 1 (s4 − s3 ) (17.86)

The same result can be achieved much more easily by a total balance of the entire
component, see Fig. 17.13b. Due to the internal heat exchange with a temperature
difference between the fluids greater than zero, entropy is generated even if both
flows are free of dissipation, i.e. no pressure drop occurs.

19 Generation of entropy in system C is caused by heat transfer, not by dissipation.


442 17 Components and Thermodynamic Cycles

17.2 Thermodynamic Cycles

In this section, the focus is on thermodynamic cycles. Thermodynamic cycles have


already been discussed to motivate the second law of thermodynamics in Chap. 15. So
far, however, these cycles have been represented exclusively20 in black box notation,
i.e. only the inflows respectively outflows are considered. Now a technical realisation
of cyclic processes is shown in detail, i.e. a technical integration of required com-
ponents into a machine. The considered machines are operated with a working fluid
that undergoes changes of state. If this working fluid within the machines reaches its
initial state after several changes of state, this is called a thermodynamic cycle. In
such a cycle, all state values Z reach their initial value, i.e.

dZ = 0 (17.87)

A distinction is made between clockwise and counterclockwise cycles. Typical tech-


nical applications are:
• Thermal engines (“clockwise cycles”)
Closed gas turbine cycles
Steam power plants
Internal combustion engines (Otto- and Diesel)
Open gas turbine plants
• Heat pump, fridge (“counterclockwise cycles”)
Absorption heat pump/fridge
Compression heat pump/fridge

17.2.1 Carnot Process

A Carnot process is an idealised cyclic process consisting of two reversible isothermal


and two reversible adiabatic changes of state21 :
• (1) → (2):
Isothermal expansion
• (2) → (3):
Reversible, adiabatic expansion
• (3) → (4):
Isothermal compression
• (4) → (1):
Reversible, adiabatic compression

20 An exception is the Carnot process, which has already been discussed in detail.
21 Since each individual change of state is reversible, the entire process is also reversible.
17.2 Thermodynamic Cycles 443

Fig. 17.14 Carnot process

Carnot’s cycle has already been introduced in Sect. 15.3.1 to predict the maximum
work that can be gained from heat at a given temperature T > Tenv . Note that a Carnot
process achieves the highest possible efficiency at a given temperature potential.
Anyhow, the process is illustrated in a p, v- and a T, s-diagram, see Fig. 17.14. In
this case the process is operated as a clockwise cycle, i.e. as a thermal engine. The
enclosed area in the p, v diagram indicates the work that can effectively be gained.22
In contrast, the enclosed area in the T, s-diagram represents the net heat, since the
process runs without dissipation.23 Furthermore, in a thermodynamic cycle, the first
law of thermodynamics obeys, cf. Eq. 11.14,
 
Q+ W = 0. (17.88)

Thus, the enclosed area in the T, s-diagram is identical with the effective work.

17.2.2 Joule Process

The Joule process is typically an ideal reference cycle for gas turbines, respectively
jet engines, i.e. clockwise cycles. The working fluid is treated as an ideal gas. The
fluid is therefore not subject to any phase change. The required components and the
process layout are shown in Fig. 17.15. However, the process can also be operated
counterclockwise. Figure 17.16 shows the clockwise cycle in a p, v- and a T, s-
diagram. It consists of the following changes of state:

• (1) → (2): Adiabatic, reversible compression


In a compressor, the fluid is isentropically compressed, i.e. work is supplied. The
pressure and temperature of the fluid increase while the specific volume decreases.

22This is due to the fact that no dissipation occurs and outer energies are ignored.
23Note that the area beneath a change of state in a T, s-diagram is the summation of specific heat
and specific dissipation.
444 17 Components and Thermodynamic Cycles

Fig. 17.15 Joule process—layout clockwise cycle

Fig. 17.16 Joule process

• (2) → (3): Isobaric heat transfer


Since the heat exchanger has no pressure losses, the change of state of the fluid
must be reversible.24 The supplied heat causes an increase of entropy, as can be
seen in the T, s-diagram.
• (3) → (4): Adiabatic, reversible expansion
The fluid expands isentropically in a turbine, i.e. work is released. Pressure and
temperature decrease, the specific volume increases.
• (4) → (1): Isobaric heat transfer
Heat is released into the environment. Since the change of state occurs at constant

24 As long as potential and kinetic energies are ignored. Nevertheless, the entire heat exchanger
generates entropy because a temperature gradient is required for heat transfer. However, the T, s-
diagram only shows the fluid being circulated. Heat is supplied to the fluid, regardless of where the
heat comes from. The only cause of dissipation in the fluid is a (mechanical) pressure loss. Note
that the entropy generation during heat transfer occurs at the interface of the two systems, i.e. at the
imperfection in the wall caused by T > 0.
17.2 Thermodynamic Cycles 445

Fig. 17.17 Clausius


Rankine process

pressure, there is no dissipation.25 Entropy needs to decrease due to the release of


heat.
The enclosed areas in the p, v- and T, s-diagrams represent the effective specific
work released by the process. Note that the entire Joule process is free of dissipation,
so the enclosed area in the T, s-diagram represents only the effective specific heat.
According to the first law of thermodynamics, see Eq. 17.88, it is identical to the
effective specific work.

17.2.3 Clausius Rankine Process

The Clausius-Rankine process is the reference cycle for an ideal steam power plant.
In contrast to the Joule process, the operating fluid is water, which is subject to a
phase change. However, the basic principle of the Clausius-Rankine process and the
Joule process is the same. The layout of the Clausius-Rankine cycle is shown in
Fig. 17.17. Since the liquid changes its state of aggregation, it cannot be treated as
an ideal gas. Therefore, the process is examined in detail in Part II, see Sect. 22.1.1.
However, the most important aspects are briefly presented. Figure 17.18 shows a
corresponding p, v- and T, s-diagram. In contrast to ideal gases, both diagrams are
characterised by limiting, bell-shaped curves that separate the liquid/wet vapour/gas
range. The following changes of state run in a Clausius Rankine process:
• (1) → (2): Isentropic compression
The liquid fluid is compressed in a pump. Since the specific volume of liquids is
small compared to gases, the required specific pressure work

2
y12 = v dp (17.89)
1

is small as well.

25 As long, as potential and kinetic energies are ignored.


446 17 Components and Thermodynamic Cycles

Fig. 17.18 Clausius Rankine process

• (2) → (3): Isobaric heat transfer + phase change


The heat supply is supposed to be without pressure losses and thus reversible, as
the change of potential and kinetic energy is neglected, i.e.

3
wt = 0 = v d p +ψ23 + ea,23 . (17.90)
 

2 =0
 

=0

Hence, it is
ψ23 = 0. (17.91)

Due to the heating, the temperature of the liquid rises26 first. As soon as vaporisa-
tion begins, the temperature is constant and all the thermal energy is used for the
phase change.27 Once the phase change is complete, the temperature continues to
rise.
• (3) → (4): Isentropic expansion
The pressurised superheated steam now drives a steam turbine and expands. In
the process, technical work is released and the pressure of the fluid decreases.
The idealised Clausius-Rankine process assumes that the expansion is adia-
batic/reversible, i.e. isentropic.
• (4) → (1): Isobaric heat transfer + phase change
To reach the initial state (1), the fluid must finally be cooled down. This is achieved
by a cooler/condenser. Since the change of state is free of pressure losses, it must
be reversible under the same conditions as in (2) → (3). The entire condenser,
i.e. fluid/enclosure/environment, is irreversible because a temperature gradient
is required for heat transfer. However, the emphasis is on the fluid, so that the

26 This is called sensible heat.


27 This is called latent heat.
17.2 Thermodynamic Cycles 447

generation of entropy can only be due to pressure losses, refer to Sect. 17.1.5 for
details.

17.2.4 Seiliger Process

The Seiliger process is an ideal reference cycle for internal combustion engines,
which is presented briefly. For further details refer to [25]. The fluid is supposed to
be an ideal gas, i.e. no phase changes take place. In addition, the following premises
are made:
• The cylinder chamber is always completely filled with ideal gas.
• The compression ratio  is the same as in the real engine, i.e.

Total volume before compression Vswing + Vcomp


= = (17.92)
Rest volume after compression Vcomp

with π 2
Vswing = d s. (17.93)
4
Typical average compression ratio for Otto engines is typically 8 . . . 11, for Diesel
engines it is in the range of 15.6 . . . 21.6, see [1].
• The mass in the cylinder stays constant, i.e. the system is closed.
• The combustion process is replaced by heating. The amount of heat is equal to the
heat released during combustion.
• All walls are to be adiabatic.
In reality, a distinction must be made between two different combustion concepts:
Ideal Otto engines assume that combustion, i.e. heat supply, takes place at constant
volume. In contrast, ideal Diesel engines are characterised by a constant pressure
during combustion. Both special processes, the equal-volume and equal-pressure
processes, cannot be achieved in reality. However, the Seiliger process is a mixed
reference cycle that takes both principles into account and is a good approximation
to real Otto/Diesel engines. Figure 17.19 shows p, v- and T, s-diagrams for all three
processes, i.e. Otto, Diesel and Seiliger processes. The changes of state are as follows:
• (1) → (2): Isentropic compression
Work is supplied, the specific volume decreases while the pressure rises.
• (2) → (2’): Isochoric heat transfer
The first part of the combustion, i.e. at constant volume corresponding to the
ideal Otto engine, is conducted. The pressure continues to rise. The combustion is
replaced by supply of heat.
• (2’) → (3): Isobaric heat transfer
The second part of the combustion, i.e. at constant pressure corresponding to the
448 17 Components and Thermodynamic Cycles

Fig. 17.19 Seiliger process

ideal Diesel engine, is conducted. The specific volume increases so that work is
released. Again, combustion is replaced by the supply of heat.
• (3) → (4): Isentropic expansion
The gas expands, i.e. expansion work is released. The specific volume increases,
the pressure decreases.
• (4) → (1): Isochoric heat transfer
According to the second law of thermodynamics, heat is released to reach the
initial state (1). Heat transfer replaces the release of gas.

17.2.5 Stirling Process

The functional principle of a Stirling process can best be explained with Fig. 17.20. In
contrast to conventional cylinder-piston engines, the Stirling engine shown consists
of two pistons that can move in parallel, i.e. the volume between the two pistons is
constant, or that can move against each other, so that the volume can decrease or
increase. The volume delimited by the two pistons is filled with a gas. In addition,
there is a regenerator, i.e. a thermal accumulator, inside the gas volume. Such a
regenerator can be easily realised, e.g. by a metal mesh which is characterised by a
high thermal conductivity. The entire process is shown in a p, v- and T, s-diagram,
see Fig. 17.21: The Stirling process is an idealised process, i.e. all changes of state
are reversible. For the exchanged work, this means that if the mechanical energies
are neglected,28 the following partial energy equation applies:

w = wV + wmech + ψ = wV , (17.94)
 


=0 =0

28 Potential energies do not play a role in a horizontal arrangement. The system is considered to be
at rest at the end of each change of state, i.e. kinetic energies are not relevant.
17.2 Thermodynamic Cycles 449

Fig. 17.20 Stirling process—possible technical principle

Fig. 17.21 Stirling process— p, v- and T, s-diagram


450 17 Components and Thermodynamic Cycles

i.e. the works consist only of volume works.


• (1) → (2): Isothermal compression
The left piston starts to move to the right while the right piston is fixed. Thereby,
the fluid is compressed. To keep the temperature constant, heat is released to the
environment:
w12 + q12 = u 2 − u 1 = 0 (17.95)



>0 <0

The regenerator is assumed to be warm.29 Cooling leads to a decrease in entropy.


The specific volume decreases, the pressure increases.
• (2) → (3): Isochoric heating
To keep the volume constant, the two pistons move parallel to the right. Con-
sequently, the cold gas must pass through the warm regenerator so that heat is
transferred from the regenerator to the gas. This is an internal heat transfer that
cools the regenerator. The heating of the gas is the cause of the increase in entropy.
The volume is fixed, so the temperature rises.

In order to conduct this change of state reversibly, the internal heat transfer between
regenerator and working fluid must also be reversible. This means that the heat
must be transferred with a T → 0, cf. Sect. 14.2. The same applies to the second
internal heat transfer at the change of state (4) → (1).
• (3) → (4): Isothermal expansion
The gas expands as the right piston moves further to the right, while the left piston
remains fixed. To keep the temperature level constant, heat is supplied from the
environment:
w34 + q34 = u 4 − u 3 = 0 (17.96)



<0 >0

The supply of heat leads to an increase in entropy. The specific volume increases,
the pressure decreases.
• (4) → (1): Isochoric cooling
The last change of state is required to reach the initial state (1) and thus close the
cycle. The hot gas passes the cold regenerator by moving the two pistons in parallel
to the left. This internal heat transfer cools the gas and heats the regenerator, i.e.
the thermal accumulator is charged. The heat release causes a decrease in entropy.
The volume is fixed so that pressure and temperature decrease.
It is obvious that heat is transferred with every change of state. However, a distinction
must be made between internal heat transfer and external heat transfer, i.e. between
the system and the environment. For the determination of the thermal efficiency,
only what is exchanged with the environment across the system boundary is relevant.
Stirling’s basic idea is that the internal heat supply is balanced by the internal heat
release, i.e. without additional thermal interaction from the outside: For the isochoric

29 As will be seen at the end of the cycle.


17.2 Thermodynamic Cycles 451

change of state (2) → (3) the first law of thermodynamics obeys

q23 + w23 = u 3 − u 2 = cv (T3 − T2 ). (17.97)




=0

For the change of state (4) → (1) it follows accordingly, i.e.

q41 + w41 = u 1 − u 4 = cv (T1 − T4 ). (17.98)




=0

Since both changes of state are isochoric, there is no work transferred. Consequently,
it follows that
q23 = −q41 (17.99)

since
T3 = T4 (17.100)

respectively
T1 = T2 . (17.101)

Thus, the thermal energy stored within the regenerator from (4) → (1) is identical
to the thermal energy extracted from the regenerator from (2) → (3).

Let us now focus on the heat transfer with the environment in order to derive the
thermal efficiency of the Stirling cycle. The specific heat supplied q results from the
first law of thermodynamics applied to the isothermal change of state (3) → (4), i.e.

w34 + q34 = u 4 − u 3 = 0. (17.102)

Hence, the specific heat supply for an ideal gas is, cf. Fig. 13.13,
 
v4
q = q34 = −w34 = RT ln > 0. (17.103)
v3

The specific heat release q0 to the environment can be derived from the first law of
thermodynamics from (1) → (2), i.e.

w12 + q12 = u 2 − u 1 = 0. (17.104)

Thus, the specific heat release for an ideal gas is, see Fig. 13.13,
 
v2
q0 = q12 = −w12 = RT0 ln < 0. (17.105)
v1
452 17 Components and Thermodynamic Cycles

According to thermal engines, the gained work w can be calculated from released
and supplied heat, cf. Fig. 15.3. The first law of thermodynamics for thermodynamic
cycles states that the energy into the machine is balanced by the energy out of
the machine. Thereby, only the exchange of energy across the system boundary is
relevant. Thus, it is30
q = |q0 | + w. (17.106)

The thermal efficiency is the ratio of benefit and effort transferred across the system
boundary, i.e.
w q − |q0 | |q0 |
ηth = = =1− . (17.107)
q q q

Substitution of the specific heats results in


    
 RT0 ln v2  RT ln v1
|q0 | v1 0 v2
ηth = 1 − =1−   =1−  . (17.108)
q RT ln vv43 RT ln vv43

Since v1 = v4 and v2 = v3 , see Fig. 17.21, it follows that

T0
ηth = 1 − = ηC (17.109)
T

The perfect Stirling process achieves Carnot efficiency, i.e. the highest possible effi-
ciency at a given temperature potential. Any irreversibility, e.g. internal friction,
reduces the efficiency. Furthermore, the best efficiency can only be achieved if the
internal storage (regenerator) can be charged and discharged perfectly, i.e. the inter-
nal heat transfer works with T → 0. In reality, however, this cannot be achieved
because a temperature potential T > 0 is required to transfer heat.

17.2.6 Compression Heat Pump

So far, clockwise cycles have been discussed. Now the focus is on heat pumps, i.e. a
counterclockwise cycles. However, these cycles are more complex because they are
usually operated with real fluids that are subject to a phase change. For this reason,
these cycles are dealt with in detail in Part II. To complete Part I, the basic principle
of a compression heat pump is briefly described, see Fig. 17.22. The purpose of a heat
pump is to use heat from the environment for heating, see Sect. 15.2.1, where a heat

30The notation of the first law of thermodynamics applied here states that the sign is taken into
account by balancing the inputs and outputs. This requires that each energy is counted as an absolute
value. Thus, the heat release q0 , which actually is negative, must be taken as an absolute positive
value but accounted for on the outgoing energy side.
17.2 Thermodynamic Cycles 453

Fig. 17.22 Compression heat pump—layout

pump has been presented as a black box. In doing so, the power required for heating,
i.e. the effort, is reduced, as part of the heating demand is covered by the utilised
environmental heat. Ambient heat is supplied in the vaporiser: The fluid is operated
at low pressure so that vaporisation takes place at low temperatures. This allows heat
from the cold environment to be transferred to the fluid. The vaporised fluid is then
compressed so that pressure and temperature rise. Heat is released in the condenser,
e.g. to a water circuit for heating, so that the vaporised fluid cools isobarically in an
idealised manner. This heat release causes a phase change so that the fluid leaves the
condenser as a liquid. To close the thermodynamic cycle, the fluid is finally throttled
to the lower pressure so that the cycle can start again. In the following, each change
of state is summarised and shown in a T, s-diagram, see Fig. 17.23. For real fluids,
i.e. fluids that can change their state of aggregation, the T, s-diagram is divided into
three regions by a limiting, bell-shaped curve:
1. Liquid, left from the limiting curve
2. Wet steam,31 within the enclosed area
3. Steam, right from the limiting curve
The T, s-diagram for real fluids is further discussed in Part II. However, the changes
of state in a compression heat pump are:

31 Liquid and vapour occur at the same time.


454 17 Components and Thermodynamic Cycles

Fig. 17.23 Compression heat pump: sketch and illustration in a T, s-diagram

• (1) → (2)
The pressure of the fluid is increased by a compressor. The specific technical work
required is wt . In the T, s-diagram, this change of state is assumed to be adiabatic
and reversible, i.e. isentropic.
• (2) → (3)
Heat is released in a condenser, e.g. to a room to be heated. This occurs at a
high temperature, respectively a high pressure. The fluid finally condenses while
releasing heat isobarically. The specific heat released is supposed to be qout .
• (3) → (4)
Adiabatic throttling causes the fluid to expand to the lower operating pressure
level. While the specific enthalpy remains constant, the entropy must increase.
This is because the internal friction within the adiabatic throttle is the cause of the
required pressure loss.32
• (4) → (1)
By external heat transfer from the environment, e.g. air, water, soil, a fluid is
vaporised at low temperature, respectively low pressure. In the best case, this
occurs without pressure loss and is therefore reversible.33 The supplied specific
heat is qin .

17.2.7 Process Overview

Table 17.1 gives an overview of the thermodynamic cycles discussed so far. It shows
the scheme of changes of state and illustrations in p, v- and T, s-diagrams. It also

32 In case kinetic and potential energies are ignored.


33 In case kinetic and potential energies are ignored.
17.2 Thermodynamic Cycles 455

Table 17.1 Overview thermodynamic cycles

shows how the process can be operated: Left means cycles counterclockwise, i.e.
heat pump/refrigerators. Right means clockwise cycles, i.e. thermal engines.

Problem 17.1 A compressor unit consists of an adiabatic low pressure compressor


(ηsC,1 = 0.91), an intercooler and an adiabatic high pressure compressor (ηsC,2 =
0.91) in series connection. Air, that can be treated as an ideal gas with R = 0.287 kgkJK
and c p = 1.004 kgkJK , is compressed in the low-pressure compressor from p1 =
1.0 bar and ϑ1 = 15 ◦ C to p2 = 6.1 bar. After that, the air is cooled down in the
interstage cooler to ϑ3 = 25 ◦ C while the pressure sinks to p3 = 5.9 bar. Finally, the
air is compressed in the high-pressure compressor to p4 = 35 bar. The volume flow
3
of the air intake of the low-pressure compressor is V̇1 = 10000 mh .
(a) Sketch the layout of the plant and illustrate the changes of state schematically
in a p, v-diagram and in a T, s-diagram.
(b) What is the power consumption Pt12 of the low-pressure compressor?
(c) What is the heat flux Q̇ 23 released in the intercooler?
(d) What is the power consumption Pt34 of the high-pressure compressor?

Solution

(a) The layout of the system is shown in Fig. 17.24. All three components are con-
nected in series. The process is depicted in both a p, v-diagram and a T, s-
diagram, cf. Fig. 17.25.
456 17 Components and Thermodynamic Cycles

Fig. 17.24 Sketch of the layout to Problem 17.1

Fig. 17.25 p, v- and T, s-diagram to Problem 17.1

(b) In order to calculate the power consumption Pt12 of the low-pressure compressor,
the first law of thermodynamics is applied,34 i.e.

Q̇ 12 +Pt12 = ṁ (h 2 − h 1 ) = ṁc p (T2 − T1 ) . (17.110)




=0

The mass flow rate follows the thermal equation of state

p1 V̇1 = ṁ RT1 . (17.111)

Note that in steady state the mass flux ṁ is constant, whereas the volume flux is
not. Solving for the mass flux leads to

p1 V̇1 kg
ṁ = = 3.3589 . (17.112)
RT1 s

34 Since there is no information about potential and kinetic energy, both are neglected. Relevant
information for kinetic energy could be, for example, a volume flow in combination with the cross-
section of a pipe.
17.2 Thermodynamic Cycles 457

The other unknown is the temperature T2 after the low-pressure compressor.


This cannot be calculated directly, but the isentropic efficiency must be applied.
In a first step, a perfect adiabatic, i.e. isentropic, compressor is examined. The
fictitious change of state (1) → (2s) is therefore isentropic, i.e.
  κ−1
κ
p2
T2s = T1 . (17.113)
p1

Note that this fictitious compressor is also working within the pressure range
p1 . . . p2 . The isentropic exponent is
cp
κ= = 1.4003. (17.114)
cp − R

In a next step, the temperature T2 results from the definition of the isentropic
efficiency
h s T2s − T1
ηsC,1 = = . (17.115)
h T2 − T1

The rearrangement results in

T2s − T1
T2 = T1 + = 502.47 K. (17.116)
ηsC,1

Hence, the power consumption follows Eq. 17.110, i.e.

Pt12 = ṁc p (T2 − T1 ) = 722.76 kW (17.117)

(c) The heat flux Q̇ 23 obeys the first law of thermodynamics for the heat exchanger,
i.e.
Q̇ 23 + Pt23 = ṁ (h 3 − h 2 ) . (17.118)


=0

The application of the caloric equation of state results in

Q̇ 23 = ṁc p (T3 − T2 ) = −689.03 kW (17.119)

(d) To calculate the power consumption Pt34 of the high-pressure compressor, the
first law of thermodynamics is applied in combination with the isentropic effi-
ciency according to (b), i.e.

Q̇ 34 +Pt34 = ṁ (h 4 − h 3 ) = ṁc p (T4 − T3 ) . (17.120)




=0
458 17 Components and Thermodynamic Cycles

First, the fictitious isentropic change of state (3) → (4s) needs to be calculated,
i.e.   κ−1
p4 κ
T4s = T3 = 495.97 K. (17.121)
p3

Thus, temperature T4 follows:

T4s − T1
T4 = T3 + = 515.53 K. (17.122)
ηsC,2

Second, the power consumption results in

Pt34 = ṁc p (T4 − T3 ) = 733.09 kW (17.123)

Alternatively, an overall balance including the entire plant (1) → (4) can be
conducted. The first law of thermodynamics for this case is

Q̇ 34 + Pt12 + Pt34 = ṁ (h 4 − h 1 ) = ṁc p (T4 − T1 ) . (17.124)

Hence, the power consumption of the high-pressure compressor is

Pt34 = ṁc p (T4 − T1 ) − Q̇ 34 − Pt12 = 733.09 kW. (17.125)

Problem 17.2 An open heat pump working with air, see Fig. 17.26, is built up of an
adiabatic compressor (ηsC = 0.89), a heat exchanger, in which the compressed air
releases heat to a water circuit, and an adiabatic turbine (ηsT = 0.89). Compressor
and turbine are mechanically connected. The air is sucked out of the environment
with p1 = penv = 1 bar, ϑ1 = ϑenv = 0 ◦ C and compressed to p2 = 4.2 bar. In the
heat exchanger the air is cooled down to ϑ3 = 85 ◦ C while the pressure decreases
to p3 = 4.05 bar. In the turbine the air expands to p4 = penv = 1 bar. Air can be
treated as an ideal gas with c p = 1.004 kgkJK = const. and R = 0.287 kgkJK . Within
the heat exchanger the water is heated up from ϑWe = 70 ◦ C to ϑWa = 90 ◦ C. The
specific heat capacity of water is cW = 4.19 kgkJK = const. Water can be treated as an
incompressible liquid.
(a) Calculate the temperature T2 .
(b) A heat flux of Q̇ H = 1 MW is supplied to the water in the heat exchanger. What
are the mass flux ṁ of the air and the mass flux ṁ W of the water?
(c) Sketch the temperature profiles of the two fluid flows in the heat exchanger as a
function of the specific enthalpy of the air. How is the heat exchanger operated?
(d) Calculate the power consumption P12 of the compressor and the released power
P34 of the turbine. What power P is effectively required in order to run the heat
pump? What is the coefficient of performance ε = Q̇PH of the heat pump?
17.2 Thermodynamic Cycles 459

Fig. 17.26 Sketch to


Problem 17.2

(e) Calculate the flux of exergy Ė x,H , that the water absorbs, as well as the exergetic
efficiency εex = ĖPx,H of the open heat pump.

Solution

(a) In order to determine temperature T2 the fictitious isentropic change of state


(1) → (2s) is calculated first, see Fig. 17.27a, i.e.
  κ−1
κ
p2
T2s = T1 = 411.68 K (17.126)
p1

The definition of the isentropic efficiency is

h s T2s − T1
ηsC = = . (17.127)
h T2 − T1

Thus, temperature T2 is

T2s − T1
T2 = T1 + = 428.8 K (17.128)
ηsC

(b) Now the heat exchanger is investigated, see Fig. 17.27b. The energy balance for
the water obeys35
ṁ W h We + Q̇ H = ṁ W h Wa . (17.129)

Thus, the mass flow rate of the water is

35 Energy in is balanced by energy out in steady state.


460 17 Components and Thermodynamic Cycles

Q̇ H
ṁ W = . (17.130)
h Wa − h We

The caloric equation of state for water, i.e. for an incompressible liquid, is

h Wa − h We = cW (TWa − TWe ) + vW ( pWa − pWe ) . (17.131)


 

=0

Furthermore, there is no information, that a pressure loss within the heat


exchanger on the water’s side occurs. This leads to the mass flow rate ṁ W

Q̇ H kg
ṁ W = = 11.93 (17.132)
cW (TWa − TWe ) s

The mass flow rate of air follows accordingly by a balance for the air, see
Fig. 17.27b, i.e.
ṁh 2 = Q̇ H + ṁh 3 . (17.133)

Hence, the mass flow rate is

Q̇ H Q̇ H kg
ṁ = = = 14.0973 (17.134)
h2 − h3 c p (T2 − T3 ) s

(c) This question is intended to clarify how the heat exchanger is to be operated, i.e.
in parallel flow or counterflow. First, the enthalpy of the air is plotted against its
temperature. The temperature is highest at the air inlet and thus the enthalpy is
also highest, since enthalpy and temperature are proportional as long as the spe-
cific heat capacity is constant. In the diagram, the air inlet and outlet are therefore
marked, cf. Fig. 17.28. The position on the x-axis is not relevant in this sketch,
but due to the enthalpy/temperature dependency, inlet and outlet temperature can
be connected by a straight line. Once the profile of air is fixed, we can focus on
the profile of water. In case of a counterflow configuration, see Fig. 17.28a, the
inlet for the water is on the opposite side of the air, i.e. where the enthalpy of air
is the lowest, which is on the left. At this position, regarding h, the inlet temper-
ature of water is marked. However, in this counterflow configuration the water
leaves on the right side, where the enthalpy of air is maximum. So the outlet tem-
perature of water is marked at this position. Since cW is supposed to be constant
and since there is no pressure loss on the water’s side,36 inlet and outlet tem-
perature can be connected by a straight line as well. Figure 17.28a indicates that
this configuration works, since the temperature of air is always greater than the
temperature of the water. Hence, heat is transferred all way long from air to water.

36 Hence, the enthalpy of water is purely a function of temperature.


17.2 Thermodynamic Cycles 461

Fig. 17.27 Sketch to Problem 17.2: a Compressor in a p, v-diagram, b Heat exchanger

Fig. 17.28 Sketch to Problem 17.2(c)

Now, let us investigate, what happens if the heat exchanger is operated in parallel
flow, see Fig. 17.28b. The profile of air is kept fix, whereas the profile of water
is just reversed. It is obvious, that in this configuration the two profiles intersect.
Thus, starting from the right side, the heat exchanger works for quite a while,
i.e. as long as the air temperature is higher than the temperature of the water.
Consequently, air releases heat to the water. However, the driving temperature
potential for heat transfer is becoming smaller and smaller. Once the curves
intersect, there is no driving temperature potential left, so heat transfer stops, i.e.
a thermal equilibrium has been reached. Apparently, this heat exchanger is not
capable to transfer the required thermal power of Q̇ H .
462 17 Components and Thermodynamic Cycles

(d) In order to calculate the power consumption of the adiabatic compressor P12 the
first law of thermodynamics is applied, i.e.

P12 + Q̇ 12 = ṁ (h 2 − h 1 ) . (17.135)


=0

Applying the caloric equation of state leads to

P12 = ṁc p (T2 − T1 ) = 2.2031 × 103 kW (17.136)

Accordingly, the power release of the turbine P34 can be calculated, i.e.

P34 + Q̇ 34 = ṁ (h 4 − h 3 ) . (17.137)


=0

Applying the caloric equation of state leads to

P34 = ṁc p (T4 − T3 ) . (17.138)

However, temperature T4 is still unknown, but can be calculated with the isen-
tropic efficiency of the turbine. In a first step the fictitious temperature T4s from
an isentropic change of state (3) → (4s) needs to be calculated, i.e.
  κ−1
κ
p4
T4s = T3 = 240.1154 K (17.139)
p3

Second, temperature T4 follows from the isentropic efficiency, i.e.

h T4 − T3
ηsT = = . (17.140)
h s T4s − T3

Hence, the required temperature T4 results in

T4 = T3 + ηsT (T4s − T3 ) = 253.0992 K. (17.141)

The released power of the turbine is

P34 = ṁc p (T4 − T3 ) = −1.4869 × 103 kW (17.142)

When the turbine drives the compressor, the technical effort is reduced. The
power required to operate the turbine/compressor unit is thus

P = P12 + P34 = 716.2074 kW (17.143)


17.2 Thermodynamic Cycles 463

Fig. 17.29 Sketch to Problem 17.2(e)

(e) In order to calculated the exergy that is supplied to the water within the heat
exchanger, there are two alternatives:
• Alternative 1
The supplied exergy is the difference of the exergies of the fluid flow at outlet
and inlet, see Sect. 16.2 and Fig. 17.29b, i.e.
 
Ė x,H = ṁ W ex,S,Wa − ex,S,We . (17.144)

Applying Eq. 16.41 for inlet and outlet with ignoring kinetic and potential
energies leads to

Ė x,H = ṁ W [(h Wa − h env ) − Tenv (sWa − senv )] +


(17.145)
−ṁ W [(h We − h env ) − Tenv (sWe − senv )] .

Simplification leads to

Ė x,H = ṁ W [(h Wa − h We ) − Tenv (sWa − sWe )] . (17.146)

Applying the caloric equations of state for incompressible liquids,37 yields


 
TWa
Ė x,H = ṁ W cW (TWa − TWe ) − Tenv cW ln = 226.3258 kW
TWe
(17.147)
Hence, the energetic efficiency of the plant is

Exergetic benefit Ė x,H


εex = = = 0.316 (17.148)
Exergetic effort P

37 Without pressure loss.


464 17 Components and Thermodynamic Cycles

• Alternative 2
The exergy balance of the system, see Fig. 17.29b, reads as

ṁ W ex,S,We + Ė x,Q = ṁ W ex,S,Wa +  Ė x,V . (17.149)

Since there is no pressure loss on the water’s side, the partial energy equation
is
2
wt = 0 = v d p +ψ12 + ea12 . (17.150)
 

1 =0
 

=0

Thus, there is no dissipation (ψ12 = 0). If there is no dissipation, there is


no entropy generation and no exergy loss. Consequently, the exergy balance
simplifies to  
ṁ W ex,S,Wa − ex,S,We = Ė x,Q . (17.151)
 

= Ė x,H

In other words, the exergy of the supplied heat is responsible for increasing
the exergy of the water. The exergy of heat38 obeys
 
Tenv
Ė x,Q = Q̇ H 1 − . (17.152)
Tm

The rearrangement results in

Q̇ H
Ė x,H = Ė x,Q = Q̇ H − Tenv . (17.153)
Tm


= Ṡa

However, this requires a balance of entropy,39 see Fig. 17.29a, i.e.

ṁ W sWe + Ṡa + Ṡi = ṁ W sWa . (17.154)




=0

Hence, the transferred entropy by heat is

Ṡa = ṁ W (sWa − sWe ) . (17.155)

The application of the caloric equation of state for incompressible fluids


results in

38Note that Tm is the thermodynamic mean temperature of the water.


39Entropy in is balanced by entropy out. There is no generation of entropy, since no pressure loss
occurs for the water.
17.2 Thermodynamic Cycles 465

TWa
Ṡa = ṁ W cW ln . (17.156)
TWe

Substitution in Eq. 17.153 leads to

TWa
Ė x,H = Ė x,Q = Q̇ H − Tenv ṁ W cW ln . (17.157)
TWe

Finally, the heat Q̇ H can be replaced according to part c), i.e.


 
TWa
Ė x,H = ṁ W cW (TWa − TWe ) − Tenv cW ln = 226.3258 kW
TWe
(17.158)

Furthermore, the energetic efficiency of the plant is

Exergetic benefit Ė x,H


εex = = = 0.316 (17.159)
Exergetic effort P

Problem 17.3 A tank with a volume of V = 1 m3 contains an ideal gas (cv =


717 kgJK , R = 287 kgJK .) In state (1) the pressure is p1 = 2 bar, the temperature is
T1 = 290 K. The gas within the tank needs to be cooled down to a temperature of
T2 = 250 K. The released energy shall be transferred to a huge environment with
a temperature of Tenv = 300 K. In order to realise the heat transfer out of the tank
respectively to the environment a temperature difference of T = 5 K is required at
any point of time.
(a) Sketch the required machine as a black box.
(b) What is the mass of the gas within the tank and what pressure occurs in state
(2)?
(c) How much heat needs to be released by the tank?
(d) What is the minimum work that needs to be supplied to the machine?
(e) What is the total loss of exergy E x,V for this process? What causes the loss of
exergy?

Solution

(a) The problem deals with a cooling machine. Its schematic is shown in Fig. 17.30.
(b) The mass in the tank is constant, since it is a closed system. Applying the thermal
equation of states leads to

p1 V
m= = 2.403 kg. (17.160)
RT1
466 17 Components and Thermodynamic Cycles

The pressure p2 in state (2) follows accordingly

m RT2 p1 V RT2 p1 T2
p2 = = = = 1.724 bar. (17.161)
V RT1 V T1

(c) The heat Q 0 to be released follows the first law of thermodynamics for the tank

Q 12 + W12 = U2 − U1 = mcv (T2 − T1 ) (17.162)




=0

so that
p1 V
Q 0 = Q 12 = mcv (T2 − T1 ) =cv (T2 − T1 ) = −6.8914 × 104 J.
RT1
(17.163)
(d) The minimum work W can be calculated with the first law of thermodynamics
for the cooling machine, i.e.

|Q| = |Q 0 | + W → W = |Q| − |Q 0 | . (17.164)

Obviously there are two unknowns, i.e. W and |Q|, so the second law of ther-
modynamics is applied. Since the temperature of the tank is not constant, it must
be given in differential notation, see Fig. 17.30, i.e.

|δ Q 0 | δQ
= . (17.165)
T − T Tenv + T

This equation follows the balance of entropy, i.e. entropy in is balanced by


entropy out. Since it is asked for the minimum work, the machine needs to be
reversible, i.e. there is no entropy generation. Furthermore, the balance is given
in steady state notation, i.e. no accumulation of entropy occurs. This is because
any thermodynamic cycle follows Eq. 17.87. In this case it reads as

dS = 0. (17.166)

Anyhow, with
|δ Q 0 | = −mcv dT (17.167)

the balance of entropy (17.165) simplifies to

−mcv dT δQ
= . (17.168)
T − T Tenv + T

The integration from state (1) to state (2) results in


17.2 Thermodynamic Cycles 467

Fig. 17.30 Sketch to


Problem 17.3

2 2
dT δQ Q
− mcv = = . (17.169)
T − T Tenv + T Tenv + T
1 1

Solving the integral leads to


T2
 Q
− mcv ln (T − T )  = (17.170)
T1 Tenv + T

respectively
T2 − T Q
− mcv ln = . (17.171)
T1 − T Tenv + T

Hence, Q follows

T2 − T
Q = −mcv (Tenv + T ) ln = 7.947 × 104 J. (17.172)
T1 − T

This finally results in

W = |Q| − |Q 0 | = 1.0554 × 104 J. (17.173)

(e) In order to calculate the total loss of exergy E x,V two alternatives are possible.
Both are sketched in Fig. 17.31. Note that both alternatives treat an extended
system boundary, that includes the T for the heat transfer.

• Alternative 1:

A balance of exergy reads as

E x,Q 0 @Tm + W = E x,Q@Tenv + E x,V . (17.174)


468 17 Components and Thermodynamic Cycles

Fig. 17.31 Sketch to


Problem 17.3

Since Q is transferred at ambient temperature, it follows

E x,Q@Tenv = 0. (17.175)

The loss of exergy then is

E x,V = E x,Q 0 @Tm + W. (17.176)

In order to solve this equation, the thermodynamic mean temperature of the tank
is required, i.e.

Q0 Q0 mcv (T2 − T1 ) (T2 − T1 )


Tm = = T
= T
= = 269.5 K (17.177)
Sa12 mcv ln T1
2
mcv ln T1
2
ln TT21

with40 ⎛ ⎞
⎜ T2 V2 ⎟
S = Si12 +Sa12 = m ⎜
⎝cv ln + R ln ⎟ . (17.178)

T1 V1 ⎠
=0  

=0

This leads to the loss of exergy, cf. Eq. 17.176, i.e.


 
Tenv
E x,V = E x,Q 0 @Tm + W = |Q 0 | 1 − = 2.7557 × 103 J. (17.179)
Tm

40 The partial energy equation for the tank reads as W12 = 0 = W12,V + W12,mech + 12 . Since
there is no volume work, respectively mechanical work, there is no dissipation and no entropy
generation in the tank. For dissipation to occur, the fluid would have to be in motion.
17.2 Thermodynamic Cycles 469

• Alternative 2:

A balance of entropy for the entire system, see Fig. 17.31, follows41

|δ Q 0 | δQ
+ δSi = . (17.180)
T Tenv

Substituting the heat results in

mcv dT δQ
− + δSi = . (17.181)
T Tenv

The integration leads to

T2
dT Q
− mcv + Si,12 = . (17.182)
T Tenv
T1

Hence, the entire generation of entropy is

Q T2 J
Si,12 = + mcv ln = 9.1858 . (17.183)
Tenv T1 K

The loss of exergy then is

E x,V = Si,12 Tenv = 2.7557 × 103 J. (17.184)

The loss of exergy is due to the irreversible heat transfer with T > 0.

41Again in differential notation, since the temperature in the tank is not constant. In this extended
system there is entropy generation due to the heat transfer, that is now part of the system boundary.
Part II
Real Fluids and Mixtures
Chapter 18
Single-Component Fluids

This chapter focuses on single-component fluids, i.e. pure fluids that are not mixed
with other fluids. In Part I, idealised single-component fluids have already been pre-
sented: Ideal gases, which follow the known thermal and caloric equations of state,
and incompressible fluids, whose specific volume remains constant. For incompress-
ible fluids, the caloric equations of state can be easily adapted. However, the behaviour
of real fluids differs from the idealised models. In particular, real fluids can change
their state of aggregation. Phase changes occur in many ways in nature, e.g. in humid
air1 or combustion processes.2 Therefore, this first chapter of Part II forms an essen-
tial basis for further understanding. The first goal is to describe the thermodynamic
state of real fluids.

18.1 Ideal Gas Versus Real Fluids

So far, see Part I, an ideal gas approach has been used to describe the thermodynamic
state of a system. However, ideal gases are subject to several model assumptions,
which are briefly listed:
• Large distances between the particles3

The interactions between the randomly moving particles are based on perfect
elastic collisions. The greater the specific volume, the more negligible these

1 Water vapour from the air can condense and form rainfall.
2 By mixing water into a combustion chamber, the combustion temperature can be controlled, as
the vaporisation process of the water makes an energetic contribution.
3 Particle is a synonym for molecules respectively atoms.

© Springer Nature Switzerland AG 2022 473


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_18
474 18 Single-Component Fluids

interactions are. Therefore, a gas behaves like an ideal gas at high temperatures
and low pressure: The intermolecular forces become less important compared to
the kinetic energy of the gas.

• The particles of an ideal gas have a negligibly small expansion

Particles are treated as mass points, respectively punctiform. The empty space
between the molecules is dominant compared to the size of the particles.

• Ideal gases obey the equations of state

The thermal equation of state, which was derived in Chap. 6 and which can also
be derived with the help of statistical thermodynamics, is as follows

pv = RT. (18.1)

The caloric equations of state derived in Chap. 12 are:

du = cv dT (18.2)

and
dh = c p dT (18.3)

and
cv R cp R
ds = · dT + · dv = · dT − · d p. (18.4)
T v T p

Many real gases behave like ideal gases under atmospheric conditions, i.e. at standard
temperature and pressure, with sufficient accuracy for technical applications, see [5],
e.g. nitrogen, oxygen, hydrogen, noble gases and carbon dioxide. However, at high
pressures and low temperatures,4 real gases behave as follows:
• The spatial expansion of the particles and the interacting forces cannot be
neglected.
• The thermal equation of state can no longer be applied. The deviation from the
thermal equation of state for ideal gases is thus taken into account, for example,
by the following modification, i.e.

pv = Z RT (18.5)

Z is a dimensionless compressibility factor equal to one for ideal gases. Figure 18.1
shows the compressibility factor for air. Obviously, for real gases, such a com-
pressibility factor need not be constant. In general, Z depends on pressure and
temperature.

4 Hence, at small specific volumes.


18.1 Ideal Gas Versus Real Fluids 475

Fig. 18.1 Compressibility factor Z for air according to [4]

Fig. 18.2 Changes of aggregation state

• In addition, at high pressures and low temperatures, liquefaction of real fluids is


possible, i.e. a change in the state of aggregation occurs.

Therefore, real fluids can occur in different states of aggregation. In the solid state,
the binding forces between the particles form solid structures. By supplying energy, a
fluid state can be achieved, i.e. the binding forces are loosened by thermally induced
movement of the particles. Further energy supply can bring the fluid into the gaseous
state so that the kinetic energy of the particles is increased. Possible changes in the
state of aggregation are shown in Fig. 18.2. The transition from solid to liquid is
denoted as freezing and the reverse as melting. When a liquid turns into gas, it is
called vaporisation. The way back is condensation. Sublimation is a change of state
476 18 Single-Component Fluids

in which the solid phase passes directly into the gaseous phase without being liquid
at any time. This occurs for instance when dry ice,5 e.g. carbon dioxide, below the
triple point6 is supplied with heat: The dry ice sublimates, i.e. it leaves the solid
aggregate state and becomes gaseous directly.
The same can be observed when water is in a solid state7 in a cold and dry
atmosphere, e.g. in winter: The partial pressure of water in the atmosphere, which is
explained in Chap. 19, is rather low, so that when heat is supplied, the phase change
takes place below the triple point of water.8 This is the reason why water sublimates
under these conditions. It is therefore more efficient to hang washed laundry outside
in winter than to dry it indoors, where the air may already be saturated with water
vapour.
However, these two examples show what is covered in Part II: First, it must be
clarified how a phase change of real fluids can be quantified. Secondly, a change of
state (solid ↔ liquid, liquid ↔ gaseous, solid ↔ gaseous) leads to sudden changes in
the physical properties of a real fluid, which must be accounted for by adapted equa-
tions of state. Thirdly, thermodynamic approaches are required to treat the mixing
of fluids, e.g. dry air and vapour.

18.2 Phase Change Real Fluids

18.2.1 Example: Isobaric Vaporisation

Let us assume that a closed system in state (1) is filled with a liquid, e.g. water, see
Fig. 18.3. A freely movable piston closes the system. The pressure of the system can
therefore be easily adjusted, since the mass of the piston and the ambient pressure
keep the pressure in the cylinder in thermodynamic equilibrium, i.e. all forces acting
on the system are balanced. It is further assumed that there is no other substance
in the system, so that it is free of any air, for example. After the supply of thermal
energy, a new state of equilibrium is reached. The pressure is the same as before
because ambient pressure and the mass of the piston are assumed to be constant.
However, the temperature increases. If the supply of heat continues, a state (2) is
reached where boiling begins, i.e. the fluid starts to vaporise. At this point, the first
vapour bubble appears and begins to rise in the fluid. The temperature at which
vaporisation starts depends on the fluid and its pressure, e.g. water at 1 bar starts
boiling at almost 100 ◦ C. During vaporisation, the temperature remains constant
despite the further supply of heat. Obviously, the supplied heat Q does not trigger

5 This does not contain any water.


6 This is explained later in Sect. 18.2.3. For carbon dioxide, however, the temperature at the triple
point is −56.5 ◦ C.
7 Thus, ice.
8 The triple point of water is at 0.01 ◦ C.
18.2 Phase Change Real Fluids 477

Fig. 18.3 Isobaric vaporisation—schematic illustration

a temperature increase, but causes a change in the state of aggregation.9 State (3) in
Fig. 18.3 shows the system vaporising: although its temperature and pressure remain
constant, its volume and thus its specific volume increase. The energy supplied is
needed for the phase change and not for a further increase in temperature. The state (3)
also exists in a thermodynamic equilibrium. However, a vapour ratio x is defined10 :
mass vapour mV
x= = (18.6)
total mass mV + mL

By definition, the vapour ratio in-between states (2) and (4) is in the range11 of 0 . . . 1.
Once vaporisation is complete, i.e. the last liquid has completely turned to vapour,
see state (4) in Fig. 18.3, the temperature of the fluid continues to rise as additional
heat is supplied, see state (5).
The area between the states (2), i.e. the start of boiling, and (4), i.e. the completion
of the phase change, is referred to as the wet steam region, since dry saturated steam
and boiling liquid are present simultaneously.
In state (5), the steam is overheated/superheated because its temperature is greater
than the temperature of the wet steam at the corresponding pressure, i.e. greater than
the temperature at which boiling begins.

9 This is named latent heat, as the supplied heat does not cause a measurable caloric effect, i.e.
rise of temperature, but a phase change that can not be detected by a thermocouple for instance. In
contrast, supplied heat that causes a temperature rise, is called sensible heat.
10 Unfortunately, x has several meanings in Part II.
11 For states (1) and (5) it does not make sense to define a vapour ratio x, since either pure liquid or

pure vapour is present.


478 18 Single-Component Fluids

Theorem 18.1 In vaporisation, the temperature of the fluid depends on its pressure,
i.e. isobaric vaporisation is also isothermal.
Consequently, a state in the wet steam region cannot be described unambiguously
only by pressure p and temperature T , as has been the case for ideal gases, see also
Sect. 5.1.2.

Vaporisation of a real fluid

It is known from Part I that a p, v-diagram is indispensable in thermodynamics, as


it illustrates e.g. the process value work. Therefore, the task now is to derive such a
p, v-diagram for a real fluid, e.g. a vaporising liquid as shown in Fig. 18.3. In a simple
experiment, isobaric vaporisation is now carried out and all thermal state values are
measured in order to plot them in a p, v-diagram.
In a first step, the states (1) to (5) are depicted in a p, v-diagram, see Fig. 18.4a.
The pressure for such a change of state is fixed and shall be pA so that all states lie
on a horizontal straight line. States (1) and (2) are close to each other because the
increase in specific volume is rather small when a liquid is heated. State (3) is in
the wet steam region, so the specific volume increases rapidly due to the increasing
vapour ratio. Experience shows that vapour occupies a larger volume than a liquid.
In a next step, the pressure is increased to pB respectively to pC , e.g. by increasing
the mass of the piston. The measured states (1) to (5) are again plotted in the p, v-
diagram, see Fig. 18.4b, c. As before, state (2) indicates the boiling point and (4) the
end of the phase change. When conducting the experiment, the temperature has also
been carefully determined so that the isotherms can be illustrated in the p, v-diagram
as well, cf. Fig. 18.4d. Note that the upper isotherms in the p, v-diagram indicate a
greater temperature than the isotherms at a lower position. Thus, at any pressure, state
(1) has a lower temperature than state (2). In the wet steam region, the temperature
is constant as long as vaporisation takes place at constant pressure, see Theorem
18.1. However, compared to an ideal gas, see e.g. Fig. 7.2, the isotherms resemble
a hyperbola which is interrupted for the wet steam region and forms a horizontal
plateau in this range. Obviously, this region is a discontinuity compared to an ideal
gas. As already mentioned, pressure and temperature do not allow a unambiguous
indication of the thermodynamic state in the wet steam region.12
Now all initial boiling points, i.e. states (2), and all states that have just finished the
phase change, i.e. states (4), are connected, see Fig. 18.4e. The left curve represents
x = 0 and is called the saturated liquid line. The right curve is represented by x = 1
and is called the saturated vapour line. The final p, v-diagram is shown in Fig. 18.4f.
Both the saturated liquid line (x = 0) and the saturated vapour line (x = 1) meet
at the so-called critical point (cp), which is fluid-specific and can be identified by pcp
and Tcp . If the temperature of the fluid is greater than Tcp , the fluid can no longer be

12 Pressure and temperature form a value pair. For each pair of values, however, vapour contents x
between 0 and 1 exist in the wet vapour region, so that the volume occupied can vary greatly, i.e. v
is not unambiguously fixed.
18.2 Phase Change Real Fluids 479

Fig. 18.4 Isobaric vaporisation—schematic illustration in a p, v-diagram


480 18 Single-Component Fluids

liquefied, no matter how great its pressure. In contrast, any vapour with a temperature
below the critical temperature Tcp can be liquefied by increasing the pressure of the
fluid. The region between the bell-shaped saturated liquid/vapour line is the wet
steam range, i.e. the two-phase range. In this region, the temperature is exclusively
a function of the pressure, see Theorem 18.1. To the left of the x = 0 line the fluid
is in the liquid state, to the right of the x = 1 line the fluid is in the gaseous/vapour
state. By definition, steams/vapours are gaseous fluids whose thermodynamic state
is close to the liquefaction point. In contrast, extremely superheated vapours are
denoted gases.

18.2.2 The p, v, T -state Space

In the previous section, the phase change from liquid to vapour has been explained.
However, an isobaric phase transition from liquid to solid state and vice versa shows
a similar behaviour, i.e. the temperature remains constant. For example, when water
is cooled at 1 bar,13 it reaches 0 ◦ C. At this point, the phase transition from liquid to
solid begins and the first ice crystals are formed. While further cooling down, the
temperature remains constant at 0 ◦ C, but the ratio of ice to liquid increases. As soon
as the liquid has completely transformed into ice, i.e. the phase change is complete,
the temperature begins to drop if heat is further released.
In a so-called p, v, T -state space, the thermal state values can be easily visualised,
see Fig. 18.5. For an ideal gas, as treated in Part I, the state space shows a smooth
curve, since
RT
p = f (v, T ) = . (18.7)
v
This graph is shown in Fig. 18.5b. From a thermodynamic point of view, it has no
imperfections or discontinuities, i.e. the state surface is continuously differentiable.
For real fluids subject to phase change, the p, v, T -space, see Fig. 18.5a, shows
similar behaviour, but discontinuities are observed at each phase change. A p, v, T -
diagram for real fluids illustrates the following characteristics:
• For small pressures p and small densities ρ, i.e. large specific volumes v, the ideal
gas equation pv = RT can be applied approximately for real gases, since in this
range Fig. 18.5a, b show similar behaviour.
• When approaching the saturation vapour line, which is shown as curve b in
Fig. 18.5a, the deviations from the ideal gas law increase, as a comparison of
the two diagrams shows.
• Region 1 represents the solid phase. This range is characterised by a relatively
small specific volume that remains almost constant when temperature T and pres-
sure p are varied.

13 However, pressure does not play a major role for water, see Sect. 18.2.3.
18.2 Phase Change Real Fluids 481

Fig. 18.5 The p, v, T -diagram: a Real fluid, b Ideal gas

• Liquid state is shown in region 2. However, liquid state and solid state are linked
by region 6, i.e. the melting region. The transition from the solid to the liquid state
and vice versa is isothermal at constant pressure, while the specific volume varies
slightly.
• Region 3 shows the gas phase.
• Region 4 in the diagram represents the wet steam region which has been discussed
previously. Liquid and vapour occur simultaneously in this region, see Sect. 18.2.1.
• Region 5 is the sublimation region, i.e. solid and vapour exist at the same time.
• Region 2/3 represents the so-called supercritical gas/liquid region, i.e. T > Tcp
and p > pcp . It is not possible to distinguish between liquid and gaseous states:
the viscosity corresponds to that of a gas, the density to that of a liquid.
The limiting curves in the p, v, T -state space have been partially discussed earlier
and are denoted as follows:
• Curve a is the saturated liquid line, i.e. x = 0. It separates liquid from wet steam.
• Curve b is the saturated liquid line, i.e. x = 1. It is the limiting line between wet
steam region and gas phase.
482 18 Single-Component Fluids

• Curve c is the solidification line, that separates melting region 6 and liquid region
2.
• Curve d is the melting line, that separates solid from melting region.
• Curve e represents the triple point line connecting sublimation and wet steam
region. Along this line, the fluid can exist in all three states.
• Curve f is the sublimation line connecting sublimation region and solid region.
• Curve g is the desublimation line connecting sublimation region and gas region.
• The specified reference point is needed later when phase changes are investigated
energetically.

Examples

Isobaric heating from (1a) → (1f)

In state (1a), a fluid is in a supercooled solid state.14 Heating takes place at almost
constant specific volume until state (1b) is reached. At this point, melting begins
so that the temperature remains constant15 while the state of the fluid changes from
solid to liquid. However, the specific volume increases. The melting process is fin-
ished as soon as state (1c) is reached. The further supply of thermal energy leads to
an increase in temperature.16 The specific volume also increases slightly. The liquid
is supercooled17 until state (1d) is reached. At this point, vaporisation begins at a
constant temperature. Vaporisation is accompanied by a sharp increase in specific
volume. Vaporisation is completed in state (1e). The further supply of thermal energy
leads to overheating of the vapour, i.e. temperature and specific volume increase.

Isobaric heating from (2a) → (2b)

A so-called supercritical change of state occurs when p > pcp . Initially, the fluid
is in a liquid state. Due to heating, i.e. supply of thermal energy, the fluid becomes
gaseous without passing a wet steam region. In the intermediate states (2a) and (2b)
the fluid appears milky without clear local phase boundaries, i.e. bubbles or droplets.
It is not possible to distinguish whether the fluid is liquid or gaseous.

Isobaric heating from (3a) → (3d)

Starting from state (3a), in which the fluid is in a solid state at a low pressure, heating
takes place at almost constant specific volume. When state (3b) is reached, the fluid

14 Thus, its temperature is smaller than the temperature at which the phase change solid/liquid takes
place.
15 The thermal energy is called latent heat.
16 The thermal energy is called sensible heat.
17 Thus, its temperature is smaller than the temperature at which the phase change liquid/gas takes

place.
18.2 Phase Change Real Fluids 483

Fig. 18.6 The p, v, T -diagram: examples

begins to sublimate, i.e. it passes from the liquid to the gaseous state at constant
temperature without being liquid in between. This change of state is accompanied
by a strong increase in specific volume. In state (3c), sublimation is complete. Further
supply of heat leads to a superheated gas (3d), while the specific volume increases
continuously.

Anomaly of water

Figure 18.6 shows another interesting aspect. Unlike most other fluids, the phase
transition of water from liquid to solid state, e.g. freezing from state (1c) to state
(1b), is accompanied by an increase in specific volume, i.e. a decrease in density.
This is shown in the detailed view of Fig. 18.6. This is the reason why ice floats on
the surface of liquid water, since

ρice < ρliquid . (18.8)

Furthermore, a beer bottle bursts in the freezer because the specific volume increases
during the freezing process. In contrast, most other fluids show the characteristic
illustrated in Fig. 18.6, i.e. the change of state from (1c) to (1b) leads to a decrease
484 18 Single-Component Fluids

Fig. 18.7 Possible projections of the p, v, T -state space

in the specific volume. This anomaly can be explained by hydrogen bonding. Due to
this hydrogen bonding, the structure in the solid state requires a larger space than in
the liquid state. In addition, water has its highest density at about 4 ◦ C.

Projections of the p, v, T -diagram

Obviously, the p, v, T -diagram provides a great amount of information, so it makes


sense to bundle this information. Once the behaviour of the fluid, as shown in
Fig. 18.5, is available, the three-dimensional state spaces can be projected to produce
two-dimensional diagrams. This is clarified in the following. Possible projections
are shown in Fig. 18.7. The derived p, v-diagram has already been discussed, see
Sect. 18.2.1. It is used, as known from Part I, to determine the process value work for
thermodynamic changes of state. In addition, a p, T -diagram can be developed by a
18.2 Phase Change Real Fluids 485

Fig. 18.8 The p, T -diagram

second projection. With the help of this diagram, the state of aggregation of a fluid
can be determined at a certain pressure p and a certain temperature T . However, the
wet steam region is hidden in this diagram because the curves of the saturated liquid
and the saturated vapour are congruent in this projection. It also includes the congru-
ent solidification/melting curves and the sublimation curve. A correlation between
pressure and temperature during vaporisation, i.e. the vapour pressure curve, can
be easily established. Finally, a T, v-diagram can be derived as shown in Fig. 18.7.
These three diagrams are discussed in the following sections.

18.2.3 p, T -Diagram

In a p, T -diagram, the three different single-phase regions are separated by three


phase limiting curves. These curves meet at the so-called triple point (TP).18 At
this point, at which temperature and pressure of three phases are in thermodynamic
equilibrium, a solid, a liquid and a vapour phase can coexist. The phase limiting
curves are designated as follows:
• Vapour pressure curve, separating liquid and steam
• Melting pressure curve, separating liquid and solid
• Sublimation pressure curve, separating solid and steam
Points (1) . . . (5) indicated in Fig. 18.8 refer to the example given in Sect. 18.2.1.
Obviously, the states (2), (3) and (4), which limit the wet steam region or lie in
this region, respectively, cannot be distinguished in this p, T -diagram, since the wet
steam information is lost due to the projection. Figure 18.9 shows p, T -diagrams
qualitatively for water (a) and non-water fluids (b). Obviously, the melting pressure

18 Due to the projection, the triple point line is only perceived as a triple point.
486 18 Single-Component Fluids

Fig. 18.9 The p, T -diagram: a Water, b Non-water fluid

Table 18.1 Triple/critical point of water


State value Triple point Critical point
p in [ bar] 0.006117 220.64
ϑ in [ ◦ C] 0.01 373.95

curve is qualitatively different. In the case of water, the phase transition from liquid
to solid is almost independent of the pressure, i.e. it takes place at about 0 ◦ C. Triple
point as well as critical point have already been explained. For water, these two states
are listed in Table 18.1.

18.2.4 T, v-Diagram

A T, v-diagram is shown in Fig. 18.10. Again, the limiting curves x = 0, i.e. the line
of the saturated liquid, and x = 1, i.e. the line of the saturated vapour, meet at the
critical point (cp). As mentioned above, the critical point is fluid-specific. The region
below the saturated liquid/vapour curve is the wet steam region already introduced,
i.e. the two-phase region in which liquid and steam occur simultaneously. In the wet
steam region, temperature is exclusively a function of pressure. To the left of the
wet steam region and below the critical temperature is the liquid region. The isobars
are close together and converge with the line of the saturated liquid, i.e. x = 0.
When heated at p > pcp , the liquid changes its state of aggregation directly into the
gaseous state. Figure 18.10 includes the example given in Sect. 18.2.1 as indicated
by the states (1) . . . (5). Since this example is an isobaric vaporisation, the states are
18.2 Phase Change Real Fluids 487

Fig. 18.10 The


T, v-diagram

on the corresponding isobar. Due to the isobaric phase change, states (2), (3) and (4)
are isothermal in the wet steam region as well.

18.2.5 p, v-Diagram

The p, v-diagram has already been introduced in Sect. 18.2.1 when isobaric vapori-
sation has been discussed. However, Fig. 18.11 now shows a p, v-diagram for water.
Obviously, in the liquid region, i.e. to the left of the x = 0 curve, the specific vol-
ume changes very little even with large pressure variations. This has been the reason
why liquids have been treated as incompressible in Part I. Since the specific volume
increases greatly with vaporisation, it makes sense to use a logarithmic scale for
the specific volume v. The curve for saturated liquid, i.e. x = 0, and the curve for
saturated vapour, i.e. x = 1, enclose the wet steam region in which the fluid consists
simultaneously of boiling liquid and saturated vapour. The ratio of vapour to total
mass has been previously defined according to Eq. 18.6. The curves for a constant
vapour ratio x are also shown in Fig. 18.11. To the right of the curve x = 1 the
gaseous phase is indicated. The isotherms start at the upper left and lead to the lower
right, similar to the isotherms, i.e. hyperbolas, of ideal gases. In real fluids, however,
the curves have a constant plateau in the wet steam region due to the phase change,
where temperature is a function of pressure.
The Figs. 18.12 and 18.13 show as an example an isothermal expansion of a closed
system. In state (1), let the water be a supercooled liquid at a pressure of 200 bar
and a temperature of 100 ◦ C. This state can be fixed in a p, v-diagram, cf. Fig. 18.13.
However, the change of state follows the 100 ◦ C isotherm. During the change of
state, the pressure is now quasi statically reduced, e.g. by reducing the weight of
the piston. In addition, the temperature remains constant. This leads to state (2) and
finally to state (3), i.e. the saturated liquid state, where vaporisation begins. For water
488 18 Single-Component Fluids

Fig. 18.11 The p, v-diagram for water

Fig. 18.12 Closed system—Isothermal expansion

at 100 ◦ C this is approximately at 1 bar. While the increase in specific volume from
(1) to (2) to (3) has been very small, the increase in specific volume upon vaporisation
is large. The process shall end in state (4) when vaporisation is complete, i.e. x = 1 is
reached. One benefit of the p, v diagram is that the specific work can be visualised.
18.2 Phase Change Real Fluids 489

Fig. 18.13 Isothermal expansion (1) to (4) in a p, v-diagram

Let us assume that the process is reversible and the change in potential energy is
negligible, so that the total work from (1) to (4) is pure volume work, i.e.

4
w14 = wV,14 + wmech,14 + ψ14 = wV,14 = − p dv < 0. (18.9)
1

Obviously, the process is an expansion, as the volume increases. Work is released


and is therefore negative. The first law of thermodynamics from state (1) to state (4)
obeys19
W14 + Q 14 = U4 − U1 . (18.10)

In specific notation it is
w14 + q14 = u 4 − u 1 . (18.11)

Applying the steam table of water20 with the given states according to Fig. 18.12
results in
kJ
w14 + q14 = u 4 − u 1 = 496.3 . (18.12)
kg

Hence, the specific heat is

19 As mentioned before, outer energies shall be ignored.


20 The steam table is introduced in Sect. 18.3.7.
490 18 Single-Component Fluids

kJ
q14 = u 4 − u 1 − w14 = 496.3 − w14 > 0. (18.13)
kg 
<0

For isothermal expansion, heat must therefore be supplied to the system. In contrast
to ideal gases, the internal energy of the fluid does not remain constant, although
the process is supposed to be isothermal. However, this is due to the phase change.
Obviously, according to Fig. 18.13, the specific volume of water in the liquid range
changes only very slightly with large pressure variations, see phase change from state
(1) to state (2) and (3), respectively. These three states are on an almost perpendicular
straight line. Therefore, liquids are often treated as approximately incompressible
fluids, as has been the case in Part I.

18.2.6 State Description Within the Wet Steam Region

Section 18.2.1 has shown that pressure and temperature are not independent of each
other in the wet steam region. Thus, the thermodynamic state cannot be unambigu-
ously characterised by pressure p and temperature T , see also Sect. 5.1.2 as well
as Fig. 5.1. To identify the state in the wet steam region, the vapour ratio x, see
Eq. 18.6, has been introduced, which represents the ratio of the mass of the vapour
to the total mass. Each extensive state value Z in the wet steam region is composed
of the properties of the saturated vapour21 (x = 1) and the boiling/saturated liquid
(x = 0), i.e.
Z = ZV + ZL. (18.14)

Introducing the specific state values according to Eq. 3.13 results in

m total z = m V z V + m L z L . (18.15)

The mass of the liquid m L can be substituted by

m L = m total − m V (18.16)

so that it is
m total z = m V z V + (m total − m V ) z L . (18.17)

A division by the total mass m total leads to

z = x z V + (1 − x) z L (18.18)

z V is the specific state value of the saturated vapour state, i.e.

21 This is also referred to as dry saturated steam.


18.2 Phase Change Real Fluids 491

z V = z (x = 1) = z  . (18.19)

z L is the specific state value of the saturated liquid state,22 i.e.

z L = z (x = 0) = z  . (18.20)

Hence, the specific state values in the wet steam region, according to Eq. 18.18, obey

V
v= = xv + (1 − x)v (18.21)
m

U
u= = xu  + (1 − x)u  (18.22)
m

H
h= = xh  + (1 − x)h  (18.23)
m

S
s= = xs  + (1 − x)s  (18.24)
m

Note that ρ = xρ  + (1 − x)ρ  since ρ = 1v .

The p, v-diagram in Fig. 18.14 shows state (3) in the wet steam region, corre-
sponding to the example in Sect. 18.2.1. As already explained, the total volume is
composed of the volume of dry saturated vapour and the volume of saturated liquid.
The vapour ratio is
mV
x= . (18.25)
mV + mL

Consequently, the liquid ratio is


mL
(1 − x) = . (18.26)
mV + mL

Combing these two equations results in

x mV m 
= = . (18.27)
1−x mL m

Solving Eq. 18.21 for the vapour ratio leads to

v − v
x= . (18.28)
v − v

22 The liquid is boiling.


492 18 Single-Component Fluids

Fig. 18.14 Lever rule of the quantities

Hence, it is
v − v
1−x = . (18.29)
v − v

Equations 18.28 and 18.29 lead to

x v − v a
=  = . (18.30)
1−x v −v b

The distances a and b are given in Fig. 18.14. A combination of Eqs. 18.27 and 18.30
results in the so-called lever rule, i.e.

a m 
=  (18.31)
b m

In other words, the more dry saturated steam the state contains, the greater a. Con-
sequently, state (3) shifts to the right. If the steam is reduced, i.e. the liquid content
is increased, b increases and state (3) moves to the left.
18.3 State Values of Real Fluids 493

18.3 State Values of Real Fluids

For thermodynamic calculation and thus for the design of technical components and
machines, the previously described state diagrams are required. Obviously, complex
measurements must be carried out to determine the properties of a fluid. These prop-
erties are the basis for any thermodynamic calculation, e.g. for the first and second
law of thermodynamics. Section 18.2.1 has shown that the thermal state values, i.e.
pressure p, temperature T and specific volume v, can be easily measured. However,
to reduce the number of experiments, only a few grid points are investigated. The
values between these grid points are derived by equations of state:
• Ideal gases follow the thermal equation of state pv = RT .
• Real fluids show deviations from pv = RT , see Sect. 18.1, so the thermal equation
of state requires adjustments. Some of the possible adjustments are briefly listed
in Table 18.2.
Unfortunately, caloric state values, such as specific entropy s, specific enthalpy h
and specific internal energy u, cannot be measured and must therefore be determined
by calculation.

18.3.1 Van der Waals Equation of State

Van der Waals,23 see [26], suggested the following equation of state in order to
describe real fluids:
 
a
p + 2 (vM − b) = RM T. (18.38)
vM

Table 18.2 Further equations of state for real fluids


Name Correlation
RM T
Redlich-Kwong p= vM −b − T 0.5 vM (vM +b)
a
(18.32)
2
0.42748RM Tcp2.5
a= pcp (18.33)
RM T
Peng Robinson p= vM −b − v2 +2bvM −b2

(18.34)
M
RM T
Berthelot p= vM −b − T v2
a
(18.35)
M
a
RT − vM RT
Dieterici p= vM −b e (18.36)
p
Virial RM T =
B(T ) C(T ) D(T )
vM + v2 + + + ··· (18.37)
1
3
vM 4
vM
M

23 Johannes Diderik van der Waals (23 November 1837 in Leiden, 8 March 1923 in Amsterdam).
494 18 Single-Component Fluids

With a = 0 and b = 0 this equation passes over to the thermal equation of state for
ideal gases, see Eq. 6.25.
• b is the so-called co-volume. It describes the reduction in volume due to the finite
dimensions of the molecules. This reduces the space for the free movement of the
molecules.
• va2 is the so-called cohesion pressure. It takes into account the decrease in pressure
M
towards limiting walls due to intermolecular attraction.
Rearranging Eq. 18.38 results in

RM T a
p = p (vM ) = − 2 . (18.39)
vM − b vM

The question is how to derive the coefficients a as well as b for a real fluid: At
the critical point, the isotherm in a p, v-diagram shows an inflection point with a
horizontal tangent, see e.g. Fig. 18.5. Above the critical temperature, the isotherms
approach regular hyperbolas. Below the critical temperature, the curves according
to Eq. 18.39 have a minimum and a maximum, see Fig. 18.15, which shows a p, v-
diagram24 for CO2 . Therefore, there must be an inflection point as well. However, let
us now focus on the isothermal Tcp : For an inflection point with a horizontal tangent,
the following mathematical conditions must be fulfilled:
 
∂p
=0 (18.40)
∂vM
T cp

and  
∂2 p
=0 (18.41)
∂vM2
T cp

with
RM Tcp a
pcp = − 2 . (18.42)
vM,cp − b vM,cp

The first derivative of Eq. 18.39 obeys


 
∂p
= −
RM Tcp + 2a = 0. (18.43)
∂vM 2 3
T cp vM,cp − b vM,cp

and the second derivative follows


 2 
∂ p
=
2RM Tcp − 6a = 0. (18.44)
∂vM T cp
2
vM,cp − b
3 4
vM,cp

24 This graph has been calculated according to Eq. 18.39.


18.3 State Values of Real Fluids 495

Rearranging Eqs. 18.43 and 18.44 leads to

2a RM Tcp
3
=
2 (18.45)
vM,cp vM,cp − b

respectively
6a 2RM Tcp
4
=
3 . (18.46)
vM,cp vM,cp − b

Dividing Eqs. 18.45 and 18.46 results in

1 1

vM,cp = vM,cp − b . (18.47)
3 2
Hence, the co-volume b can be calculated, in case the critical point is known, by

1
b= vM,cp . (18.48)
3

Coefficient a follows from Eq. 18.45 with b being replaced by Eq. 18.48, i.e.

9
a= RM Tcp vM,cp . (18.49)
8
Rearranging results in
8
RM Tcp = . (18.50)
9vM,cp

The critical pressure follows from Eq. 18.42, i.e.

3RM Tcp a
pcp = − 2 . (18.51)
2vM,cp vM,cp

Substitution of RM Tcp by Eq. 18.50 brings

a
pcp = 2
. (18.52)
3vM,cp

In other words, the cohesion pressure a obeys

a = 3 pcp vM,cp
2
(18.53)
496 18 Single-Component Fluids

Fig. 18.15 Van der Waals approach for CO2

Figure 18.15 shows the van der Waals approach for CO2 . According to Eq. 18.39,
the p, v curves can be calculated for different isotherms. Obviously, the curves agree
with the expected characteristic outside the wet steam region, as e.g. a comparison
with Fig. 18.3 shows. However, the wet steam region cannot be covered by the van
der Waals equation because the isotherms in a p, v diagram do not show a horizontal
line but have maxima and minima. A fitting is done according to Maxwell,25 in order
to determine the bell-shaped curve of the wet steam region. The Maxwell approach
satisfies
C

 

p vM − vM = p dvM . (18.54)
A

This rule can be justified by the fact that the area under the curve of a change of state
in a p, v-diagram corresponds to the specific volume work. Thus, the work from A to
B should be equal to the work from B to C, i.e. the two hatched areas in the detailed
representation of Fig. 18.15 must have the same absolute value. Following this idea,
the bell-shaped curves26 can be constructed both for saturated liquid, i.e. x = 0, and
for vapour, i.e. x = 1. Under this premise, the calculated specific volume work from

25 James Clerk Maxwell (13 June 1831 in Edinburgh, 5 November 1879 in Cambridge).
26 Illustrated as dashed line.
18.3 State Values of Real Fluids 497

A to C agrees with the actual specific volume work for an isobaric phase change
from A to C.

18.3.2 Redlich-Kwong

Redlich and Kwong, see [27], suggested the following correlation for real fluids, i.e.

RM T a
p= − 0.5 (18.55)
vM − b T vM (vM + b)

with
2 2.5
0.42748RM Tcp
a= (18.56)
pcp

and
0, 08664RM Tcp
b= . (18.57)
pcp

However, this correlation, which is similar to the Van der Waals approach, is more
accurate above critical temperature.

18.3.3 Peng-Robinson

An improvement of the van der Waals equation was developed by Peng and Robinson
in 1976, cf. [28]. Their approach follows
RM T aα
p= − 2 (18.58)
vM − b vM + 2bvM − b2

with
0.457235RM Tcp2
a= (18.59)
pcp

and
0.077796RM Tcp
b= . (18.60)
pcp

α can be calculated depending on the acentric factor27 ω. For ω < 0.49 it is




2
α = 1 + 0.37464 + 1.54226ω − 0.26992ω2 1 − Tr0.5 (18.61)

27 The acentric factor is a measure of the non-sphericity (centricity) of molecules.


498 18 Single-Component Fluids

and for ω ≥ 0.49 it is



2
α = 1 + (0.379642 + (1.48503 − (1.164423 − 1.016666ω) ω) ω) 1 − Tr0.5 .
(18.62)
Tr is the reduced temperature, that follows

T
Tr = . (18.63)
Tcp

18.3.4 Berthelot

Berthelot, cf. [29], suggested the following equation

RM T a
p= − (18.64)
vM − b 2
T vM

with
a = 3Tcp pcp vM,cp
2
(18.65)

and vM,cp
b= . (18.66)
3
However, this correlation is rarely used.

18.3.5 Dieterici

According to Dieterici, see [30], the following equation of state can be applied for
real fluids, i.e.
RT − v aRT
p= e M . (18.67)
vM − b

The parameters a and b follow


vM,cp
a = 4RM Tcp (18.68)
2
and vM,cp
b= . (18.69)
2
18.3 State Values of Real Fluids 499

18.3.6 Virial Equations

Virial equations, cf. [31], are extensions of the thermal equation of state for ideal
gases and can be derived directly from statistical mechanics, i.e.

p 1 B(T ) C(T ) D(T )


= + 2 + 3 + 4 + ··· (18.70)
RM T vM vM vM vM

They include series expansions of v1M . In case the series expansion is terminated after
the first term, the virial equation turns into the thermal equation of state for ideal
gases, see [5] for details.

18.3.7 Steam Tables

Instead of calculating state values by equations of state, so-called steam tables are
usually used in technical thermodynamics, see Tables A.1, A.2, A.3, A.4, A.5, A.6,
A.7, A.8, A.9, A.10, A.11 and A.12. These tables contain exemplary thermal and
caloric state values for water28 : Tables A.1, A.2, A.3 and A.4 cover the limiting
bell-shaped curve of the wet steam region by listing state values for the saturated
liquid state ( ), i.e. x = 0, and the saturated steam state ( ), i.e. x = 1. According to
Sect. 18.2.1, pressure and temperature depend on each other, so that in addition the
vapour ratio x unambiguously identifies each specific state value

Z
z= (18.71)
m
within the wet steam region, see Sect. 18.2.6, i.e.

z = x z  + (1 − x) z  . (18.72)

Note that this equation cannot be applied to the density ρ but to the specific volume
v = ρ1 = mV , since
Z
z= . (18.73)
m
In addition to the wet steam region, the state values specific volume v, specific
enthalpy h, specific entropy s and specific heat capacity c p of water for both super-
cooled liquid and superheated steam are given in Tables A.5, A.6, A.7, A.8, A.9,

28The tables listed follow the International Association for the Properties of Water and Steam
(IAPWS). There are several digital tables available that use the IAPWS guideline, e.g. [9]. When
using state values from different sources, check whether the reference points for the caloric state
values are identical. If not, they must be corrected.
500 18 Single-Component Fluids

A.10, A.11 and A.12. Unless otherwise stated, the state values for water in this book
are taken from the steam tables and do not follow the mathematical state functions
described previously, cf. Table 18.2.
Caloric state values require a reference point (0) with which the specific internal
energy u, the specific enthalpy h as well as the specific entropy s can be compared,
see Sect. 12.4.4:
1
du = u 1 − u 0 (18.74)
0

1
dh = h 1 − h 0 (18.75)
0

1
ds = s1 − s0 . (18.76)
0

The choice of the reference state is arbitrary, but must be considered in thermody-
namic calculations:
Definition 18.1 For the listed steam tables of water29 the reference point is the
saturated liquid state (’) at the triple-point,30 i.e. at

p0 = pTP = 0.006117 bar (18.77)

and
ϑ0 = ϑTP = 0.01 ◦ C. (18.78)

The specific internal energy u is set to zero at reference level, i.e.

kJ
u0 = 0 . (18.79)
kg

Furthermore, the specific entropy s is set to zero as well, i.e.

kJ
s0 = 0 . (18.80)
kg K

Thus, the specific enthalpy h 0 at reference level follows

h 0 = u 0 + pTP v0 . (18.81)

29 For other tables, the reference point must be carefully checked.


30 Refer also to Fig. 18.5.
18.3 State Values of Real Fluids 501

Taking into account the specific volume according to the steam table, it is

N m3 kJ
h 0 = 0 + 611.66 2
· 0.0010002 = 0.00061178 . (18.82)
m kg kg

By definition, h 0 is the specific enthalpy at reference level, i.e. h 0 = h  ( pTP , ϑTP ),


see Table A.1.
Example 18.1 (State functions versus steam table) In this example, the specific
enthalpy h  of saturated boiling water at p = 1 bar is to be investigated. For real
fluids it is already known, cf. Eq. 12.99 in Sect. 12.3.2, that
   
∂v
dh = c p ( p, T ) · dT + −T + v · d p. (18.83)
∂T p

Since the specific enthalpy h  is in demand at 1 bar, Eq. 18.83 must be integrated. The
reference level for the integration is the previously introduced saturated liquid state
( ) of the triple point. Thus, both the reference level and the required state are liquid.
If the assumption from Part I is applied that the liquid is almost incompressible, the
following simplification yields  
∂v
≈ 0. (18.84)
∂T p

Hence, the equation simplifies to

dh = c p ( p, T ) · dT + v · d p. (18.85)

Since v is considered constant, integration leads to

T  p T 
h  (1 bar) − h 0 = c p dT + v dp = c p dT + v ( p − p0 ) (18.86)
T0 p0 T0

with
kJ
h 0 = h  (0.01 ◦ C) = 0.00061178 . (18.87)
kg

Thus, it applies
T 
h  (1 bar) = h 0 + c p dT + v ( p − p0 ) . (18.88)
T0

The temperature of the saturated liquid state ( ) at 1 bar is listed31 in Table A.1, i.e.

31 For this purpose, a linear interpolation must be carried out.


502 18 Single-Component Fluids

T  = 372.7559 K. (18.89)

In Table A.1 specific heat capacities are given, but these are the non-averaged, actual
values as a function of temperature. However, the required temperature-averaged
value can be calculated using the approach presented in Sect. B and Eq. 12.191,
respectively. According to Table B.6 one obtains for water:

99.6059 ◦ C kJ
c p = cp 0.01 ◦ C = 4.192 . (18.90)
kg K

The average specific volume according to Table A.1 is approximately constant, i.e.

99.6059 ◦ C 1 m3
v = v 0.01 ◦ C = 99.6059 ◦ C = 0.01 . (18.91)
ρ ◦
0.01 C
kg

This leads to

kJ
h  (1 bar) = h 0 + c p T  − T0 + v ( p − p0 ) = 418.5005 . (18.92)
kg

The calculation has shown that the determination of state values using state func-
tions is time-consuming. Furthermore, a simplification of Eq. 18.83 has been applied
with regard to compressibility. This makes the determination of state values with
steam tables much easier: In this example, the requested specific enthalpy is h  =
kJ
417.4365 kg , which follows by interpolation of Table A.1. In this table, compressibil-
ity is considered according to Eq. 18.83. In this case, however, the two approaches
differ by less than 0.3 %.
Problem 18.1 Determine the specific internal energy of saturated steam and satu-
rated liquid at a pressure of 10 bar. The working fluid is water.

Solution

In order to find u  respectively u  the steam Table A.1 can be applied. Since only
the specific enthalpies h  and h  are listed, these values must be converted for the
specific internal energy, i.e.

h  = u  + pv → u  = h  − pv (18.93)

respectively
h  = u  + pv → u  = h  − pv . (18.94)

Let us first start with the saturated liquid ( ). According to the steam table it is

kJ
h  = 762.65 (18.95)
kg
18.3 State Values of Real Fluids 503

and
1 m3
v = 
= 0.0011 . (18.96)
ρ kg

Hence, it is
kJ
u  = h  − pv = 761.5556 . (18.97)
kg

For the saturated vapour ( ) it follows accordingly

kJ
h  = 2777.1 (18.98)
kg

and
1 m3
v = 
= 0.1943 . (18.99)
ρ kg

Hence, it finally is
kJ
u  = h  − pv = 2582.8 . (18.100)
kg

Problem 18.2 8 m3 wet steam at a pressure of 9 bar have a vapour ratio of 35%.
What is the mass of the wet steam and what is its enthalpy? The working fluid is water.

Solution

The wet steam is composed of saturated liquid ( ) and saturated vapour ( ), so that

V = VL + VV (18.101)

and
m = mL + mV. (18.102)

The specific volume of the wet steam can be calculated from the ( )- and ( )-states,
since its vapour ratio x is known:

v = v (1 − x) + v x. (18.103)

Applying steam Table A.1 leads to

1 m3
v = v (9 bar) = = 0.0011211 (18.104)
ρ  (9 bar) kg

and
1 m3
v = v (9 bar) = = 0.2149 . (18.105)
ρ  (9 bar) kg
504 18 Single-Component Fluids

Thus, the specific volume is

m3
v = v (1 − x) + v x = 0.0759 . (18.106)
kg

It further is
V
v= , (18.107)
m
so that the entire mass m results in
V
m = mL + mV = = 105.3539 kg. (18.108)
v
According to
mV
x= (18.109)
m
the mass of the saturated vapour is

m V = xm = 36.8739 kg (18.110)

and the mass of the saturated liquid is

m L = m − m V = 68.48 kg. (18.111)

The entire enthalpy H of the wet steam results in

H = m L h  + m V h  . (18.112)

The specific enthalpies h  and h  are listed in the steam table, i.e.

kJ
h  = h  (9 bar) = 742.7246 (18.113)
kg

respectively
kJ
h  = h  (9 bar) = 2773.0 . (18.114)
kg

Thus, it finally follows

H = m L h  + m V h  = 153 114.4 kJ. (18.115)

Alternatively applies

H = m h  (1 − x) + h  x = 153 114.4 kJ. (18.116)
18.4 Energetic Consideration 505

Fig. 18.16 Isobaric vaporisation: a reversible, b irreversible

18.4 Energetic Consideration

18.4.1 Reversibility of Vaporisation

In a first step, the reversibility of an isobaric vaporisation ( ) → ( ), see Fig. 18.16,
is investigated. A horizontal cylinder is filled with a real fluid that undergoes a phase
transition. A freely moving, frictionless piston closes the system.
• Case (a) according to Fig. 18.16a
The first law of thermodynamics for case (a) obeys

W12 + Q 12 = U  − U  . (18.117)

The partial energy equation follows

W12 = − penv V = WV,12 + + Wmech,12 . (18.118)


  
=0

The entire work released by the system is used to compress the environment
with the moving piston.32 The mechanical work is zero because the piston moves
horizontally and the states ( ) and ( ) are supposed to be at rest. For an isobaric
change of state33 the volume work can be easily calculated,34 i.e.

32 The work W12 is transferred to the environment without loss. From the perspective of the envi-
ronment, the work W12 corresponds to the volume work of the environment. From the system’s
point of view, it is composed of the right-hand terms in Eq. 18.118.
33 Thus, the change of state is supposed to run very slowly, see Example 7.6.
34 For a quasi-static change of state the balance of forces results in p = p .
env
506 18 Single-Component Fluids

( )
WV,12 = − p dV = − penv V. (18.119)
( )

Thus, Eq. 18.118 results in


=0 (18.120)

In other words, if the vaporisation is isobaric, there is no dissipation inside the


system. Hence, the change of state as shown in Fig. 18.16a is reversible. This is
the case, if the change of state runs quasi-statically and the system is homogeneous:
Turbulence due to rising vapour bubbles, as sketched in the previous figures, does
not occur, since they cause dissipation. In such a case, however, Eq. 18.118 would
be as follows
( )
W12 = − penv V = − p dV + . (18.121)
( )

Hence, the dissipation is

( )
= − penv V + p dV > 0. (18.122)
( )

Introducing an averaged constant pressure p for the change of state ( ) to ( ) results
in
= − penv V + p V > 0 (18.123)

respectively
p > penv . (18.124)

Consequently, in the case of an expansion (V > 0), the pressure inside the sys-
tem must be p > penv if dissipation occurs.

• Case (b) according to Fig. 18.16b


Let us now have a closer look at Fig. 18.16b. Although the process is still sup-
posed to be isobaric, i.e. there is no dissipation35 within the modified system, the
choice of this new system boundary36 now causes the process to be irreversible.

35 Representing internal friction.


36 The system now includes the cylinder wall.
18.4 Energetic Consideration 507

Obviously, there is an imperfection regarding the temperature distribution, as a


temperature difference T is required for the supply of heat, see Fig. 18.16b. At
the system boundary the temperature is Tenv , inside the system it is T = Ts ( p).
This imperfection generates entropy Si,12 , see Sect. 14.2 and Fig. 14.3. According
to Fig. 18.16b, the process involving the wall is now irreversible, as the heat once
supplied is unable to leave the system against the temperature gradient. An entropy
balance obeys
S  − S  = Si,12 + Sa,12 . (18.125)

The supplied entropy due to the heat is37

Q 12,(b)
Sa,12 = . (18.126)
Tenv

Thus, it is

Q 12,(b)
Si,12 = S  − S  − >0 (18.127)
Tenv

18.4.2 Heat of Vaporisation

In a next step, the energy required for vaporisation is investigated. An isobaric heating
of a real fluid has been presented in Sect. 18.2.1. According to Fig. 18.17, the change
of state (2) → (4), i.e. from ( ) → ( ), is analysed. The first law of thermodynamics
in differential notation obeys

δq + δw = du + dea . (18.128)

With the partial energy equation

δw = − p dv + δψ + δwmech (18.129)
  
=dea

the first law of thermodynamics simplifies to

δq + δψ = du + p dv. (18.130)

Introducing the specific enthalpy

dh = du + p dv + v d p (18.131)

37Note that the temperature inside the system, where the heat passes, is relevant. According to the
boundary given in Fig. 18.16b it is Tenv .
508 18 Single-Component Fluids

Fig. 18.17 Isobaric vaporisation of water

leads to the following notation of the first law of thermodynamics

dh = δq + δψ + v d p. (18.132)

According to the second law of thermodynamics

δq δψ
ds = + (18.133)
T T
the specific heat and the specific dissipation can be substituted by

δq + δψ = T ds. (18.134)

Thus, the fundamental equation of thermodynamics results in

dh = T ds + v d p. (18.135)

For an isobaric vaporisation,38 i.e. d p = 0, it is

dh = T ds (18.136)

38As already described, the change of state is isobaric, so that no dissipation occurs if, in addition,
outer energies are neglected, cf. Example 7.6.
18.4 Energetic Consideration 509

as well as
dh = δq + δψ + v d p . (18.137)
 
=0 =0

The integration from ( ) → ( ) results in

( )
q = h  − h  = T ds. (18.138)
( )

An isobaric vaporisation from ( ) → ( ) is also isothermal, i.e. dT = 0, cf. Theorem


18.1. Thus, the specific enthalpy of vaporisation h v can be defined, i.e.


h  − h  = h v = T s  − s  (18.139)

The specific enthalpy of vaporisation h v is solely a function of temperature, i.e.

h v = h  − h  = f (Ts ) . (18.140)

Figure 18.18, which was generated using the steam table, shows this correlation. It
is obvious that the specific enthalpy of vaporisation h v decreases with increasing
temperature. For ϑ → ϑcp the specific enthalpy of vaporisation h v approaches zero.
However, due to h = u + pv it follows that

Fig. 18.18 Specific enthalpy of vaporisation h v of water


510 18 Single-Component Fluids


h v = h  − h  = u  − u  + ps v − v . (18.141)

The isobaric supply of specific enthalpy of vaporisation h v increases the internal


energy u = u  − u  , although the temperature

remains constant but the state of
aggregation changes. The second part ps v − v represents the volume work that
occurs during isobaric expansion39 of the fluid. However, according to Fig. 18.18 the
change of the internal energy dominates the volume work.

18.4.3 Caloric State Diagrams

Part I has shown that the advantage of a p, v-diagram is that specific work, i.e. a
process value, can be visualised. In addition, the T, s-diagram can be applied to
represent the other remaining process values, i.e. specific heat as well as specific
dissipation. Therefore, the focus in this section is on the relevant caloric state dia-
grams for real fluids, which show a different physical behaviour than ideal gases.
The presented diagrams can be easily constructed based on steam tables, e.g. steam
tables of water according to Tables A.1, A.2, A.3, A.4, A.5, A.6, A.7, A.8, A.9, A.10,
A.11 and A.12.

T, s-diagram

The major advantage of the T, s-diagram is that both the specific heat and the specific
dissipation can be visualised as area below the change of state. However, the T, s-
diagram for real fluids differs from the diagrams for ideal gases due to the phase
change. An example of a T, s-diagram for water is shown in Fig. 18.19. A more
detailed and enlarged T, s-diagram for water can be found in Fig. C.2. The saturation
lines, i.e. x = 0 and x = 1, form a bell-shaped curve with the critical point on its
peak. Below this bell-shaped curve is the wet steam region. To the left of the wet
steam region the liquid is supercooled, to the right of it the vapour state is located.
Isobars, i.e. p = const., in the liquid region are approximately identical to the left
saturation line (x = 0). As discussed at the beginning of this chapter, in the wet
steam region, i.e. 0 < x < 1, isobars and isotherms, i.e. T = const., are congruent.
The specific entropy s is set to zero in the liquid state of the triple point according to
the chosen reference level, see Sect. 18.3.7.
In case of an isobaric, reversible vaporisation, see Sect. 18.4.2, the specific heat
of vaporisation
h v = h  − h  = f (Ts ) (18.142)

can be illustrated as hatched, rectangular area, see Fig. 18.20, since the temperature is
constant. Obviously, the rectangular area, i.e. the specific enthalpy of vaporisation,

39 The specific volume increases largely during vaporisation and compresses the environment.
18.4 Energetic Consideration 511

Fig. 18.19 T, s-diagram of water

decreases with rising pressure respectively temperature. At the critical point the
specific enthalpy of vaporisation is approaching zero.

h, s-diagram

In addition to the T, s-diagram, an h, s-diagram is advantageous, e.g. to represent


changes of state in thermal turbomachines such as compressor or turbine. An exam-
ple of such a diagram is shown in Fig. 18.21 to characterise the regions liquid/wet
steam/vapour. Analogous to the T, s-diagram, the limiting curves x = 0 and x = 1
form a bell curve. Below the bell curve is the wet steam region, to the left of it the
supercooled liquid and to the right of it the region of the vapour. A more detailed
diagram for water is shown in Fig. C.3. Within the wet steam region, isobars and
isotherms, T = const., are congruent. Isobars, i.e. p = const., within the wet steam
region run as rising straight lines and cross the saturated steam line, i.e. x = 1,
steadily: According to the fundamental equation of thermodynamics it applies for
the isobar
dh p = T ds + v d p = T ds. (18.143)

=0

Thus, the slope of the isobar is


512 18 Single-Component Fluids

Fig. 18.20 T, s-diagram of water


dh
= T. (18.144)
ds p

In the wet steam region, the temperature remains constant during isobaric vaporisa-
tion, cf. Theorem 18.1, so the slope of the isobar must also be constant. However,
the isobars almost coincide with the x = 0 curve. Isotherms are almost horizontal
straight lines in the strongly superheated region and bend downwards when approach-
ing the saturation vapour line (x = 1). A kink is observed when crossing the x = 1
curve. Strongly superheated vapours show the behaviour of an ideal gas, i.e. the
isotherms and isenthalps run horizontally. The specific enthalpy is therefore only a
function of temperature.

log p, h-diagram

This diagram is usually used for the design of air conditioning systems. Therefore,
it is further discussed in Chap. 21. A log p, h-diagram for water is briefly shown in
Fig. 18.22, but a detailed diagram is included in the Appendix, cf. Fig. C.1.

Example 18.2 (Thermal turbo-machines) As mentioned above, a h, s-diagram is


beneficial to show changes of state in thermal turbo-machines, see Fig. 18.23. In
case these components are adiabatic, according to the first law of thermodynamics
18.4 Energetic Consideration 513

Fig. 18.21 h, s-diagram of water

Fig. 18.22 log p, h-diagram of water

for open systems in steady state40

δwt12 = dh (18.145)

40 Note that the change of kinetic as well as of potential energy is ignored. This is applied in the
following two cases.
514 18 Single-Component Fluids

the specific technical work δwt can be easily calculated. Let us have a closer look at
two examples:
1. Case 1: Adiabatic, reversible turbine, see (1) → (2) in Fig. 18.23

Superheated steam expands from pressure p1 to pressure p2 < p1 . The first law
of thermodynamics obeys

wt12 + q12 = h 2 − h 1 . (18.146)

With q12 = 0 it simplifies to

wt12 = h 2 − h 1 (18.147)

Obviously, technical work is released. The second law of thermodynamics fol-


lows
ṁ(s1 − s2 ) + Ṡi + Ṡa = 0. (18.148)

With Ṡi = 0, i.e. reversible operation, and Ṡa = 0, i.e. adiabatic change of state,
it is
s2 − s1 = 0 (18.149)

The change of state runs vertically in a h, s-diagram.

2. Case 2: Adiabatic, irreversible compressor, see (3) → (4) in Fig. 18.23

Superheated steam is compressed from pressure p3 to pressure p4 > p3 . The


first law of thermodynamics reads as

wt34 + q34 = h 4 − h 3 . (18.150)

With q34 = 0 it is
wt34 = h 4 − h 3 . (18.151)

Obviously, technical work is supplied. The second law of thermodynamics fol-


lows
ṁ(s3 − s4 ) + Ṡi + Ṡa = 0. (18.152)

With Ṡi > 0, since the change of state is supposed to be irreversible, and Ṡa = 0
it results in
s4 − s3 > 0 (18.153)

The specific entropy rises from state (3) to state (4).


18.4 Energetic Consideration 515

Fig. 18.23 Technical work in the h, s-diagram

Example 18.3 (Isobaric, isothermal, reversible change of state) In this example a


closed system, see Fig. 18.24, is investigated. The change of state (1) → (2) takes
place in the wet steam region.41 The first law of thermodynamics reads as

w12 + q12 = u 2 − u 1 . (18.154)

Following the partial energy equation, the specific work is

w12 = wV,12 + wmech,12 + ψ12 = − p (v2 − v1 ) . (18.155)


   
=0 =0

Hence, the specific heat follows according to

q12 = u 2 − u 1 + p(v2 − v1 ) = h 2 − h 1 . (18.156)

The specific enthalpies h 1 and h 2 can be calculated with the vapour ratio x, see
Sect. 18.2.6, i.e.
h 2 = (1 − x2 )h  + x2 h  (18.157)

respectively
h 1 = (1 − x1 )h  + x1 h  . (18.158)

Thus, Eq. 18.156 simplifies to

41 Changes of kinetic and potential energy shall be ignored.


516 18 Single-Component Fluids

Fig. 18.24 Isobaric, isothermal, reversible change of state

q12 = (x2 − x1 ) h v (18.159)

Alternatively, with the definition of the specific enthalpy of vaporisation, see


Eq. 18.139, it is


q12 = T s  − s  (x2 − x1 ) (18.160)

Example 18.4 (Isochoric change of state) This example focuses on a closed system
as well. A cylinder is filled with wet steam and underlies a change of state while
its volume remains constant. Hence, the change of state can be easily visualised in
a p, v-diagram, i.e. Fig. 18.25a. Furthermore, the change of state can be shown in a
T, s-diagram, see Fig. 18.25b. Obviously, two cases are conceivable:
1. Heating of wet steam from (1) → (2): The vapour ratio x increases

x2 > x1 . (18.161)

2. When heating up from state (3) with low vapour ratio x3 to state (4), the vapour
ratio x may decrease, i.e.
x4 < x3 . (18.162)

Applying the first law of thermodynamics42 results in

w12 + q12 = u 2 − u 1 . (18.163)

According to the partial energy equation, the specific work is

w12 = wV,12 + wmech,12 + ψ12 = − p (v2 − v1 ) = 0. (18.164)


   
=0 =0

42 Note that the change of kinetic as well as of potential energy is ignored.


18.4 Energetic Consideration 517

Fig. 18.25 Isochoric change of state

Hence, the specific heat follows

q12 = u 2 − u 1 = u 2 − u 1 + x2 (u 2 − u 2 ) − x1 (u 1 − u 1 ). (18.165)

Since the change of state is isochoric, the specific volume remains constant, i.e.

v1 = v2 . (18.166)

The application of the vapour ratio results in

(1 − x1 ) v1 + x1 v1 = (1 − x2 ) v2 + x2 v2 . (18.167)

Solving for the vapour ratio in state (2) yields

v1 − v1 v1 − v2


x2 = x1   +  (18.168)
v2 − v2 v2 − v2

Example 18.5 (Adiabatic change of state) An adiabatic compression from p1 to p2


of a closed system is investigated, see Fig. 18.26. The following cases are possible:
• If a reversible adiabatic, i.e. isentropic, compression is conducted at

s > scp (18.169)

the vapour ratio x increases, i.e. the vapour from (1) to (2) becomes drier. Hence,
it is
x2 > x1 . (18.170)
518 18 Single-Component Fluids

Fig. 18.26 Adiabatic change of state

• If an irreversible, i.e. si12’ > 0, adiabatic compression is conducted from (1) →


(2 ), the specific entropy rises, i.e.

s2 − s1 = sa12’ +si12’ > 0. (18.171)



=0

• If a reversible adiabatic, i.e. isentropic compression is conducted at

s < scp (18.172)

the vapour ratio decreases, i.e. the vapour from (3) to (4) becomes wetter. Hence,
it is
x4 < x3 . (18.173)

18.4.4 Clausius-Clapeyron Relation

The Clausius-Clapeyron relation correlates pressure and temperature during vapor-


isation. It can be easily derived from the following thermodynamic cycle, which is
briefly illustrated in Fig. 18.27. A real fluid in a closed system undergoes the follow-
ing reversible changes of state:
• (1) → (2): Isobaric/isothermal vaporisation
• (2) → (3): Pressure drop by d p and accordingly a temperature drop by dT
• (3) → (4): Isobaric/isothermal condensation
• (4) → (1): Pressure rise by d p and accordingly a temperature rise by dT
The p, v-diagram in Fig. 18.27a illustrates this thermodynamic cycle and can be used
to determine the volume work for each change of state. Note that the volume work
equals the projected area beneath each change of state. According to the partial
18.4 Energetic Consideration 519

Fig. 18.27 Clausius-Clapeyron relation

energy equation, cf. Eq. 11.22, the entire work equals the volume work, neglecting
both mechanical work43 and dissipation44 :
• Isobaric vaporisation (1) → (2)


w12 = − ( p + d p) v − v < 0 (18.174)

• Pressure drop by d p (2) → (3): Due to the small dv the projected area beneath
the change of state, i.e. the volume work, can be ignored compared with the work
from (1) → (2), so that
w23 = 0 (18.175)

• Isobaric condensation (3) → (4)




w34 = − p v − v > 0 (18.176)

• Pressure rise by d p (4) → (1): Due to the small dv the projected area beneath
the change of state, i.e. the volume work, can be ignored compared with the work
from (1) → (2) resp. the work from (3) → (4), so that

w41 = 0 (18.177)

Thus, the overall work, i.e. the summation of Eqs. 18.174 to 18.177 results in


w = −d p v − v . (18.178)

43 Changes of potential and kinetic energy are ignored.


44 The cycle is supposed to be reversible.
520 18 Single-Component Fluids

Fig. 18.28 Vapour pressure curve of water

The T, s-diagram in Fig. 18.27b can be applied to determine the exchanged ther-
mal energy. Since each change of state is supposed to be reversible, the specific heats
are represented by the areas beneath each change of state, i.e.
• Isobaric vaporisation (1) → (2)


q12 = (T + dT ) s  − s  (18.179)

• Temperature drop by dT (2) → (3): Due to the small ds the specific heat, i.e.
the projected area beneath (2) → (3) can be ignored compared with the projected
areas from (1) → (2) and (3) → (4), so that

q23 = 0 (18.180)

• Isobaric condensation (3) → (4)




q34 = T s  − s  (18.181)

• Temperature rise by dT (4) → (1): Due to the small ds the specific heat, i.e. the
projected area beneath (4) → (1) can be ignored compared with the projected
areas from (1) → (2) and (3) → (4), so that

q41 = 0 (18.182)

Thus, the overall heat, i.e. the summation of Eqs. 18.179 to 18.182 results in


q = dT s  − s  . (18.183)
18.4 Energetic Consideration 521

For a thermodynamic cycle it is, see Eq. 11.14


 
w=− q. (18.184)

Substituting Eqs. 18.178 and 18.183 leads to





d p v − v = dT s  − s  . (18.185)


With h v = T s  − s  it is

dp s  − s  h v
=  = (18.186)
dT v − v T (v − v )

dp
The slope dT as well as its integration, i.e. ps = f (T ), are shown in Fig. 18.28.
Possible applications of the Clausius-Clapeyron relation are:
• In a closed system, the required specific enthalpy of vaporisation h v can be
measured for a certain pressure p respectively a certain temperature

T . Further-
more, the change of the specific volume during vaporisation, i.e. v − v , can be
dp
determined. Thereby, the slope dT can be calculated according to Eq. 18.186. The
procedure can be repeated at different temperatures, so that it finally is

dp
= f (T ) . (18.187)
dT
• In a closed system, the slope of the vapour pressure curve and the associated
temperature T are measured.

If in addition the change of the specific volume
during vaporisation v − v is determined, the specific enthalpy of vaporisation
h v can be calculated according to


dp
h v = f (T ) = T v − v . (18.188)
dT

18.5 Adiabatic Throttling—Joule-Thomson Effect

This section explains the different behaviour of an ideal and a real gas at adiabatic
throttling, see also [32]. The adiabatic throttle has already been introduced as a
component in Part I: The pressure of a fluid passing a throttle is reduced by internal
friction, see Fig. 18.29, so that
p2 < p1 . (18.189)
522 18 Single-Component Fluids

Fig. 18.29 Adiabatic throttling

This can easily be achieved by reducing the cross-section of the throttle. However, this
leads to a local increase in velocity and thus to an increase in turbulence. However,
the change of state is investigated under the following premises:
• Since no technical work is exchanged across the system boundary, the system is
work-insulated, i.e. passive:
wt12 = 0. (18.190)

• The system is supposed to be adiabatic. Thus, the heat exchange to the environment
is neglected, i.e.
q12 = 0. (18.191)

• The system is in steady state, so that the equation of continuity can be applied, i.e.

ṁ 1 = ṁ 2 (18.192)

• The change of mechanical energy is to be ignored,45 i.e.

z1 = z2 (18.193)

respectively
c12 = c22 . (18.194)

Under these premises, the functioning of the throttle can be easily explained with
the partial energy equation:

2 2
wt12 = 0 = v d p + ψ12 + ea,12 = v d p + ψ12 . (18.195)
1 1

Reformulated, one obtains the following relationship between pressure change and
specific dissipation:
2
v d p = −ψ12 < 0. (18.196)
1

45 The influence of the mechanical energies has been investigated in Problem 12.5.
18.5 Adiabatic Throttling—Joule-Thomson Effect 523

The specific dissipation is therefore responsible for a pressure change d p < 0. In


addition, the first law of thermodynamics

1
2
q12 + wt12 = h 2 − h 1 + c2 − c12 + g · (z 2 − z 1 ) (18.197)
2
simplifies to
h2 − h1 = 0 (18.198)

The change of state is therefore isenthalpic. To calculate the temperature, a caloric


equation of state must be applied. According to Chap. 12, the specific enthalpy is
generally a function of pressure and temperature, i.e.

h = h ( p, T ) . (18.199)

Its total differential yields



  
∂h ∂h
dh = dT + d p. (18.200)
∂T p ∂p T
  
cp

18.5.1 Ideal Gas

Section 12.3.2 has shown, that for ideal gases it is


 
∂h
= 0. (18.201)
∂p T

The specific enthalpy is therefore solely a function of temperature, i.e.


   
∂h ∂h
dh = dT + d p = c p dT. (18.202)
∂T p ∂p T
     
cp =0

Since the change of state is isenthalpic, i.e. dh = 0, it must consequently also be


isothermal, i.e. dT = 0, for ideal gases, i.e.

T1 = T2 . (18.203)
524 18 Single-Component Fluids

18.5.2 Real Gas

For real gases, however, the enthalpy depends not only on the temperature but also on
the pressure. The change of enthalpy with respect to pressure at constant temperature
in Eq. 18.200 is described with the so-called isothermal throttling coefficient δT , i.e.
 
∂h
= δT . (18.204)
∂p T

The total differential of the specific enthalpy therefore simplifies to

dh = c p dT + δT d p. (18.205)

According to the first law of thermodynamics for an adiabatic throttling, see


Eq. 18.198, it follows
dh = 0. (18.206)

Equation 18.205 then reads as46


 
∂T δT
=− = δh (18.207)
∂p h cp

Anyway, this equation quantifies how the temperature of a real gas changes when the
pressure varies while the specific enthalpy remains constant. In contrast to ideal gases,
a temperature change can occur in real gases when there is adiabatic throttling, i.e.
when the pressure is reduced. The sign of this temperature change depends, since the
specific heat capacity c p is positive, solely on the Joule-Thomson coefficient or the so-
called isenthalpic throttling coefficient δh , which is fluid-specific. The functionality
of the adiabatic throttle is visualised in Fig. 18.30. The graph shows the isenthalps in
a T, p-diagram. While for ideal gases the curves are horizontal because the specific
enthalpy only depends on the temperature, the graphs for real gases also show a
pressure dependence. Throttling, however, means a pressure reduction, so that the
change of state moves from right to left along an isenthalp47 for an adiabatic throttle.
Obviously, three cases are possible:
• Case 1 (Temperature rises):

According to Eq. 18.207 the temperature rises, i.e. dT > 0, when the pressure
decreases, i.e. d p < 0, as long as

46The subscript h indicates that the specific enthalpy is kept constant.


47It must run along an isenthalp, since the first law of thermodynamics has shown that dh = const.
However, Fig. 18.30 shows several isenthalps—depending on the inlet state of the fluid.
18.5 Adiabatic Throttling—Joule-Thomson Effect 525

Fig. 18.30 Adiabatic


throttling (Joule-Thomson
effect)

 
∂T
= δh < 0. (18.208)
∂p h

This case is shown in Fig. 18.30 as change of state (1) → (2).

• Case 2 (Temperature drops):

According to Eq. 18.207 the temperature drops, i.e. dT < 0, when the pressure
decreases, i.e. d p < 0, as long as
 
∂T
= δh > 0. (18.209)
∂p h

This case is shown in Fig. 18.30 as change of state (1 ) → (2 ).

• Case 3 (Temperature stays constant):

According to Eq. 18.207 the temperature remains constant, i.e. dT = 0, when the
pressure decreases, i.e. d p < 0, as long as
 
∂T
= δh = 0. (18.210)
∂p h

Figure 18.31 shows Joule-Thomson coefficients for different gases according to


[33]. Joule-Thomson coefficients for air are shown in Fig. 18.32 according to [34] as
a function of pressure and temperature. The air is free of CO2 and water.

Example 18.6 In this example, the adiabatic throttle of a real gas is to be investigated
using the Van der Waals equation. To determine the temperature characteristic during
throttling, the following equation must be solved:
526 18 Single-Component Fluids

Fig. 18.31 Joule-Thomson coefficient δh for several gases

Fig. 18.32 Joule-Thomson coefficient δh for air


18.5 Adiabatic Throttling—Joule-Thomson Effect 527
   
∂T δT 1 ∂h
= δh = − =− . (18.211)
∂p h cp cp ∂p T

The left-hand side of this equation describes the change in temperature with the
change in pressure, while the enthalpy remains constant. According to the first law
48
of thermodynamics, this  isthe case with an adiabatic throttle. Following Eq. 12.98
∂h
the partial differential ∂ p can be substituted, i.e.
T

 
∂v
∂T T ∂T p
−v
= . (18.212)
∂p h cp

For a description of the thermal state values of real fluids, a physical model is required.
In this example, the Van der Waals approach is applied, see Sect. 18.3.1. However, a
Van der Waals equation is calculated with molar state values, see Eq. 18.39, so molar
state values are also applied to Eq. 18.212. With
vM
v= (18.213)
M
and
c p,M Cp
cp = = (18.214)
M M
it is
 
∂vM
∂T T ∂T p
− vM
= . (18.215)
∂p h Cp

∂vM
In order to solve ∂T p
, the Van der Waals equation is applied, i.e.

RM T a
p= − 2 . (18.216)
vM − b vM

As long as vM b this equation can be simplified mathematically to49


 
RM T b a
p= 1+ − . (18.217)
vM vM vM RM T

Rearranging results in
RM T RM T a
p= +b 2 − 2 . (18.218)
vM vM vM

48 Changes of kinetic and potential energy are ignored.


49 2 = v2 +
However, this is the first approximation: Comparing Eqs. 18.216 and 18.217 leads to vM M
b , which is true for vM b.
2
528 18 Single-Component Fluids

vM
Multiplying with p
leads to

RM T RM T a
vM = +b − . (18.219)
p vM p vM p

For the correction terms including a and b the thermal equation of state pvM = RM T
is applied,50 so that
RM T a
vM = +b− . (18.220)
p RM T

∂vM
Hence, the partial differential ∂T p
can be solved, i.e.
 
∂vM RM a
= + . (18.221)
∂T p p RM T 2

Finally, this term can be substituted in Eq. 18.215, i.e.


  RM T
+ a
− vM
∂T p RM T
= . (18.222)
∂p h Cp

In this equation vM can be replaced by Eq. 18.220, so that


   
∂T 1 2a
= δh = −b (18.223)
∂p h Cp RM T

In order to solve this equation, the variables need to be separated:

dT dp
= . (18.224)
2a
RM T
−b Cp

An integration from state (1) to state (2) leads to51

T2 p2
dT dp p2 − p1
= = . (18.225)
2a
RM T
−b Cp Cp
T1 p1

Solving results in
 
2a ln (2a − b RM T ) T T2 p2 − p1
− 2
− = . (18.226)
b RM b T1 Cp

50 In fact, applying the correlation for ideal gases is the second approximation.
51 It is assumed, that C p = const..
18.5 Adiabatic Throttling—Joule-Thomson Effect 529

Fig. 18.33 Adiabatic throttling of air

Thus, it finally is

1 2a 2a − b RM T2 p2 − p1
− (T2 − T1 ) − 2 ln = . (18.227)
b b RM 2a − b RM T1 Cp

For a throttling from p1 to p2 < p1 the outlet temperature T2 can be calculated


implicitly. However, it is much easier to solve Eq. 18.227 numerically. This can be
done, for example, for air with the following van der Waals parameters:

Pam6
a = 135.8 × 10−3 (18.228)
mol2
and
m3
b = 0.0364 × 10−3 (18.229)
mol
as well as
J
C p = 28.9554 . (18.230)
mol K
Figure 18.33a shows the temperature dependent isenthalpic throttle coefficient δh
according to Eq. 18.223. Figure 18.33b, as a result of numerical integration, visualises
the throttling of air from an initial state (1), i.e. pressure p1 = 20 bar and temperature
T1 = 300 K, to a pressure p2 = 1 bar. The graph shows that the outlet temperature
is approximately T2 = 295.19 K.
Problem 18.3 A steam vessel contains 10 t liquid water and 15 kg saturated steam
at a pressure of 10 bar. The vessel shall be treated as a closed system.
(a) Calculate the total enthalpy of the content of the vessel.
530 18 Single-Component Fluids

(b) What is the volume of the vessel?


(c) What is the mass of liquid water and steam when the pressure is reduced to
2 bar?
(d) What is the temperature in the vessel after the pressure has dropped?
(e) Sketch the change of state in a p, v- and T, s-diagram.
Solution

The vapour ratio can be calculated as follows


mV
x= = 0.001 49. (18.231)
mV + mL

(a) The extensive enthalpy H follows

H = HV + HL = m V h V + m L h L = m V h  + m L h  . (18.232)

According to the steam table it is

kJ
h  = h  (10 bar) = 762.6828 (18.233)
kg

respectively
kJ
h  = h  (10 bar) = 2777.1 . (18.234)
kg

Thus, the extensive enthalpy is

H = m V h  + m L h  = 7.6685 × 106 kJ. (18.235)

Alternatively, one can follow the approach



H = m total (1 − x) h  + xh  = 7.6685 × 106 kJ. (18.236)

(b) The volume is shared by vapour and liquid, i.e.

V = VV + VL = m V vV + m L vL = m V v + m L v . (18.237)

According to the steam table it is

m3
v = v (10 bar) = 0.0011 (18.238)
kg

respectively
m3
v = v (10 bar) = 0.1943 . (18.239)
kg
18.5 Adiabatic Throttling—Joule-Thomson Effect 531

Thus, the extensive volume is

V = m V v + m L v = 14.1876 m3 . (18.240)

Alternatively, one can follow the approach



V = m total (1 − x) v + xv = 14.1876 m3 . (18.241)

(c) The vessel is supposed to be a closed system. We further assume that the vessel
is rigid, i.e. that its volume is constant. At the new pressure p2 = 2 bar it is

V = m V v + m L v = 14.1876 m3 (18.242)

as well as
m total = m V + m L = m  + m  = 10015 kg. (18.243)

According to the steam table it is

m3
v = v (2 bar) = 0.0011 (18.244)
kg

respectively
m3
v = v (2 bar) = 0.8857 . (18.245)
kg

A combination of Eqs. 18.242 and 18.243 results in

V − v m total
m = = 10011 kg (18.246)
v − v

as well as
m  = m total − m  = 4 kg. (18.247)

Thus, the vapour ratio x has decreased, i.e.

m 
x2 = = 4.0254 × 10−4 . (18.248)
m total

Alternatively, it is
V
v= = 0.0014 = const. (18.249)
m total

Since
v = (1 − x) v + xv (18.250)
532 18 Single-Component Fluids

Fig. 18.34 Solution to Problem 18.3(e)

it follows

v − v (2 bar)
x2 = = 4.0254 × 10−4 . (18.251)
v (2 bar) − v (2 bar)

Thus, the mass of the steam results as follows

m  = x2 m total = 4 kg. (18.252)

(d) A linear interpolation in the steam table leads to a temperature ϑ2 = 120.21 ◦ C.


(e) The change of state is sketched in Fig. 18.34. According to the p, v diagram, no
volume work is transferred because the volume remains constant. No mechan-
ical work is done either, since the centre of gravity is fixed. Since no work is
exchanged, the dissipation follows from the partial energy equation, i.e.

W12 = 0 = W12,V + W12,mech + 12 . (18.253)


     
=0 =0

Hence, the change of state is free of dissipation, i.e. 12 = 0. Consequently,


the hatched area in the T, s-diagram exclusively represents the specific heat.
Obviously, the system must be cooled to achieve p2 = 2 bar.

Problem 18.4 10 kg steam with a pressure of 3 MPa and a temperature of 290 ◦ C


(state 1) are enclosed in a cylinder/piston system. By releasing heat, the steam is
isochorically cooled to 200 ◦ C (state 2). Further heat is released isothermally until a
pressure of 2.5 MPa (state 3) is reached.
(a) Illustrate the states respectively changes of state qualitatively in a p, v-diagram.
(b) Determine the volume and vapour ratio in each state.
18.5 Adiabatic Throttling—Joule-Thomson Effect 533

Fig. 18.35 Solution to


Problem 18.4

The working fluid is water.

Solution
(a) Figure 18.35 shows the change of state in a p, v-diagram qualitatively.52
• State (1):

It is known from state (1), that

p1 = 3 MPa = 30 bar (18.254)

respectively
ϑ1 = 290 ◦ C. (18.255)

The vaporisation temperature at 30 bar is 233.86 ◦ C, see steam table. Thus,


since ϑ1 > ϑs (30 bar) = 233.86 ◦ C, state (1) is overheated.

• State (2):

The system is closed, so that the mass remains constant, i.e. m 1 = m 2 . Fur-
thermore, the change of state is isochoric, i.e. V1 = V2 . Consequently, the
specific volume follows

V1 V2
v1 = = = v2 . (18.256)
m1 m2

Hence, the change of state runs vertically in a p, v-diagram. The second infor-
mation required to determine the state thermodynamically unambiguously is
the temperature

52 In order to solve this problem, the knowledge of the isotherms in a p, v-diagram is required.
534 18 Single-Component Fluids

ϑ2 = 200 ◦ C. (18.257)

Unfortunately, it is not yet clear whether state (2) is in the wet steam region
as shown in the p, v diagram. This will be proven later in part (b).

• State (3):

It is known from state (3), that

p3 = 2.5 MPa = 25 bar (18.258)

respectively, since the change of state runs isothermally,

ϑ3 = ϑ2 = 200 ◦ C. (18.259)

The vaporisation temperature according to 25 bar is 223.96 ◦ C, see steam


table. Thus, since ϑ3 < ϑs (25 bar) = 223.96 ◦ C, state (3) is supercooled.

(b) For the volume and the vapour ratio, respectively, this results in:

• State(1) → overheated steam

The steam table shows, that

m3
v1 = v (30 bar, 290 ◦ C) = 0.0792 . (18.260)
kg

For volume V1 it is
V1 = mv1 = 0.792 m3 . (18.261)

• State (2)

Since the change of state (1) → (2) is isochoric, it is V2 = V1 = 0.792 m3 ,


3
i.e. v2 = v1 = 0.0792 mkg . Now, let us examine, whether state (2) lies in the
wet steam region. According to the steam table it follows

m3 m3
v (200 ◦ C) = 0.0012 < v2 < v (200 ◦ C) = 0.1272 . (18.262)
kg kg

Thus, state (2) is located in the wet steam region. Its vapour ratio is

v2 − v (200 ◦ C)
x2 = = 0.619. (18.263)
v (200 ◦ C) − v (200 ◦ C)
18.5 Adiabatic Throttling—Joule-Thomson Effect 535

• State(3) → supercooled liquid:

The steam table shows, that

m3
v3 = v (25 bar, 200 ◦ C) = 0.00116 . (18.264)
kg

For volume V3 it is
V3 = mv3 = 0.0116 m3 . (18.265)
Chapter 19
Mixture of Ideal Gases

So far, single-component liquids have been treated, i.e. fluids with an unchanging
chemical composition. Part I of the book has focused on ideal gases and incompress-
ible fluids, which are typical representatives of single-component fluids. In Chap. 18
fluids have been investigated that undergo a phase change, i.e. a fluid can occur in
solid, liquid and gaseous state—but its chemical composition remains constant even
with a phase change. This chapter clarifies how a mixture of several gases with dif-
ferent chemical compositions can be treated thermodynamically. Gas mixtures thus
consist of two or more pure gases. Each gas is considered ideal in this chapter.1 Fur-
thermore, these gases must not react chemically2 with each other. An example of a
gas mixture is dry, atmospheric air. Its volume concentrations are as follows, see [35]:
• Nitrogen (N2 = 78.08 vol. − %)
• Oxygen (O2 = 20.95 vol. − %)
• Argon (Ar = 0.93 vol. − %)
• Carbon dioxide (CO2 = 0.04 vol. − %)
• Neon Ne, helium He, methane CH4 , krypton Kr, hydrogen H2 , dinitrogen
monoxide N2 O, xenon Xe in the ppm range3
Humid air, i.e. a mixture of dry air and water, i.e. vapour, liquid, solid, is handled in
Chap. 20.

1 At sufficiently low pressure, gas mixtures can also be considered ideal, cf. [5].
2 However, chemical reactions are treated in Chaps. 23 and 24.
3 1 ppm ≡ 1 × 10−4 vol. − %.

© Springer Nature Switzerland AG 2022 537


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_19
538 19 Mixture of Ideal Gases

Fig. 19.1 Mixture of gases

19.1 Concentration Specifications

Section 5.1.3 has shown that e.g. pressure p and temperature T in multiphase systems
are not sufficient to uniquely identify a thermodynamic system. A typical example
is the wet steam region of a real fluid, where saturated steam and boiling liquid
coexist and pressure and temperature depend on each other. Obviously, additional
information about the composition of the mixture is needed to describe the state
unambiguously. However, there are several ways to characterise the concentration
of a component i in a mixture, see Fig. 19.1:
• Mass concentration ξi

Defined as the ratio of the partial mass4 m i of a component i to the total mass
m total , i.e.
mi
ξi = . (19.1)
m total

Obviously, it is

n
m total = m 1 + m 2 + · · · + m n = mi (19.2)
i=1

so that the division by the total mass m total results in


n
mi
m1 m2 mn i=1
+ + ··· + = = 1. (19.3)
m total m total m total m total

Thus, it follows, that


n
ξ1 + ξ2 + · · · + ξn = ξi = 1 (19.4)
i=1

4 The entire mass of component i in Fig. 19.1.


19.1 Concentration Specifications 539

• Molar concentration xi

Defined as the ratio of the partial molar amount5 n i of a component i to the total
molar amount n total , i.e.
ni
xi = . (19.5)
n total

Obviously, it is

n
n total = n 1 + n 2 + · · · + n n = ni (19.6)
i=1

so that the division by the total molar amount n total results in


n
ni
n1 n2 nn i=1
+ + ··· + = = 1. (19.7)
n total n total n total n total

Thus, it follows, that


n
x1 + x2 + · · · + xn = xi = 1 (19.8)
i=1

• Volume concentration σi

Defined as the ratio of the partial volume6 Vi of a component i to the total volume
Vtotal , i.e.
Vi
σi = . (19.9)
Vtotal

Obviously, it is

n
Vtotal = V1 + V2 + · · · + Vn = Vi (19.10)
i=1

so that the division by the total volume Vtotal results in


n
Vi
V1 V2 Vn i=1
+ + ··· + = = 1. (19.11)
Vtotal Vtotal Vtotal Vtotal

5 The entire amount of particles i in Fig. 19.1.


6 The total volume of particles i in Fig. 19.1 if all particles could be separated and sorted. However,
the particles in thermodynamic equilibrium occupy the entire volume, see Sect. 19.2.
540 19 Mixture of Ideal Gases

Thus, it follows, that


n
σ1 + σ2 + · · · + σn = σi = 1 (19.12)
i=1

• Partial pressure ratio πi

Defined as the ratio of the partial pressure7 pi of a component i to the total pressure
ptotal , i.e. pi
πi = . (19.13)
ptotal

• Molarity ci

Defined as the ratio of the partial molar amount n i of a component i to the total
volume Vtotal , i.e. ni
ci = . (19.14)
Vtotal

• Volume ratio ψi

Defined as the ratio of the partial volume Vi of a component i to the partial volume
Vk of a component k, i.e.
Vi
ψi = . (19.15)
Vk

• Partial density ρi

Defined as the ratio of the partial mass m i of a component i to the total volume
Vtotal , i.e.
mi n i Mi
ρi = = . (19.16)
Vtotal Vtotal

19.2 Dalton’s Law

In a mixture of ideal gases, each component behaves as if it were alone, i.e. it occupies
the entire volume of the mixture Vtotal . Figure 19.2 shows the consequence for an
adiabatic mixture of two ideal gases: At the beginning, the two gases 1 and 2 are
separated by a separating wall, see left figure. In this state, both gases are supposed
to have the same pressure p and the same temperature T . Gas 1 has a mass m 1
and occupies a volume of V1 . For gas 2 it is m 2 and V2 respectively. Now the wall is
removed so that both gases can share8 the total volume Vtotal = V1 + V2 . According to

7 See Sect. 19.2.


8 This process is called diffusion and is based on a chemical potential.
19.2 Dalton’s Law 541

Fig. 19.2 Adiabatic mixing of two gases according to Dalton

Dalton, equilibrium is reached as soon as both gases have distributed themselves over
the entire volume.9 Obviously, the intensive state values of pressure and temperature
are not the trigger for the change of state, as they are identical for both gases in
state (A). Rather, it is the concentration differences, i.e. the chemical potential, of
the two components that drive the system into thermodynamic equilibrium (B). This
phenomenon is examined later in Example 24.4.

Temperature After the Mixing

Let us assume, that during the mixing no energy is exchanged with the environment.10
Consequently, according to the first law of thermodynamics

W + Q = 0 = U1 + U2 = m 1 cv,1 (TB − T ) + m 2 cv,2 (TB − T ) (19.17)

the temperature of the mixture is


TB = T (19.18)

The mixture therefore has the same temperature as the gases in the initial state (A).

Pressure

Now, the focus is on the pressure of state (B). Since, left and right Fig. 19.2 represent
equilibrium states (A) and (B) and both gases are supposed to be ideal, the thermal
equation of state can be applied, i.e.
• Gas 1:

Before the mixing, i.e. state (A), it is

9However, reaching this equilibrium takes a while.


10There is no heat crossing the system boundary, since the system is supposed to be adiabatic.
Furthermore, no work passes the system boundary, see Fig. 19.2.
542 19 Mixture of Ideal Gases

pV1 = m 1 R1 T. (19.19)

After the mixing, i.e. state (B), it is

p1 (V1 + V2 ) = m 1 R1 T. (19.20)

Obviously, in state (B) gas 1 occupies a larger volume than in state (A), while
temperature and mass remain constant. Thus, the gas 1 in the mixture, i.e. state
(B), has a smaller pressure11 p1 than in state (A). Combining these two equations
brings
V1
p1 = p. (19.21)
V1 + V2

• Gas 2:

Before the mixing it is


pV2 = m 2 R2 T. (19.22)

After the mixing it is


p2 (V1 + V2 ) = m 2 R2 T. (19.23)

Combining these two equations results in

V2
p2 = p. (19.24)
V1 + V2

Obviously, according to Eqs. 19.21 and 19.24, the pressures of gases 1 and 2 in the
mixture are smaller than their initial pressures p in state (A).
Theorem 19.1 Each component within a mixture has an individual pressure that is
smaller than the total pressure of the mixture. This is denoted partial pressure pi .
With the partial pressures12 according to Eqs. 19.21 and 19.24 the total pressure pB
of the mixture in state (B) follows

V1 V2
pB = p1 + p2 = p+ p. (19.25)
V1 + V2 V1 + V2

11 This is denoted partial pressure: Let us theoretically assume that a pressure sensor sensitive to
gas 1 is immersed in state (B). The pressure correlates with the number of particles impinging on
the sensor. Due to the decreasing particle number density in (B) compared to (A), the impacts of
gas 1 are reduced. This is expressed by the partial pressure p1 . The same applies accordingly to gas
2.
12 A real pressure sensor does not distinguish between the particles of the different gases—it

always shows the entire pressure. Nevertheless, the different particles have a specific ratio on the
total pressure. The partial pressures pi represent the part of the pressure caused by the component
i.
19.2 Dalton’s Law 543

Thus, it is
pB = p1 + p2 = p. (19.26)

In other words, the pressure of the mixture is equal to the initial pressures of the two
components, i.e. under the given conditions the change of state is isobaric:

pB = p (19.27)

Based on Eq. 19.26, an extension can easily be made for a mixture of n compo-
nents. Obviously, the total pressure is equal to the sum of all partial pressures:

ptotal = p = p1 + p2 + · · · + pn = pi (19.28)
i

Division by ptotal brings


p1 p2 pn
1= + + ··· + (19.29)
ptotal ptotal ptotal

For the partial pressure ratio, see Sect. 19.1, it finally is



πi = 1 (19.30)

Partial Energy Equation

It has already been shown that the temperature remains constant during the adiabatic
mixing process according to Fig. 19.2. This is due to the fact that no energy has
passed the system boundary and the two gases have already been in thermal equilib-
rium. However, both gases expand according to Fig. 19.2, so the temperature should
decrease, as discussed in Part I: Reversible expansion of an ideal gas releases work,
so heating was required to keep the temperature constant. The partial energy equa-
tion,13 see Eq. 11.22, can be applied to show how the temperature remains constant
during adiabatic mixing, i.e.

B B
W = 0 = WV,1 + WV,2 +  = − pGas1 dVGas1 − pGas2 dVGas2 +  (19.31)
A A

Since the gases are ideal, the thermal equation of state can be applied, i.e.

13 The entire work W = 0, since no energy is exchanged with the environment.


544 19 Mixture of Ideal Gases

B B
m 1 R1 T m 2 R2 T
0=− dVGas1 − dVGas2 + . (19.32)
VGas1 VGas2
A A

The change of state is isothermal, i.e.

V1 + V2 V1 + V2
 = m 1 R1 T ln + m 2 R2 T ln . (19.33)
V1 V2

In other words, by applying the thermal equation of state, see Eqs. 19.19 and 19.22:

V1 + V2 V1 + V2
 = pV1 ln + pV2 ln >0 (19.34)
V1 V2

Obviously, it is the dissipation  that keeps the temperature constant. The dissipation
is driven by the volume work of the expanding gases 1 and 2.

Entropy Generation

The application of the second law of thermodynamics leads to

SB − SA = Sa +Si . (19.35)

=0

Sa disappears, since the system is adiabatic. A further simplification can be made as


follows
B
δ
SB − SA = Si = . (19.36)
T
A

Since T = const. it is
B
δ 
SB − SA = Si = = . (19.37)
T T
A

Finally one obtains14

V1 + V2 V1 + V2
Si = m 1 R1 ln + m 2 R2 ln (19.38)
V1 V2

Respectively, with Eqs. 19.21 and 19.24, it results in

14 Gibb’s paradox states that when two identical gases mix, there is no entropy generation, i.e.
Si = 0, see Sect. 19.3.4.
19.2 Dalton’s Law 545

p p
Si = m 1 R1 ln + m 2 R2 ln (19.39)
p1 p2

The following is an extension to n components15


n
p
Si = m i Ri ln (19.40)
i
pi

19.3 Laws of Mixing

19.3.1 Concentration, Thermal State Values

Concentration

The application of the thermal equation of state for a component i in state (A), cf.
Fig. 19.2, leads to
ptotal Vi = m i Ri T. (19.41)

The thermal equation of state for a component i in state (B) brings

pi Vtotal = m i Ri T. (19.42)

Thus, its combination results in

pi Vi
πi = = = σi (19.43)
ptotal Vtotal

Hence, the partial pressure ratio πi is identical with the volume concentration σi .

Equation 19.42 is now generalised for all components and summed up, i.e.
 
pi Vtotal = m i Ri T. (19.44)
i i

The mass m i of a component can be substituted by n i Mi , i.e.


  
pi Vtotal = n i Mi R i T = n i RM T. (19.45)
i i i

15The proof is simple, since instead of 2 components n components have to be considered, starting
with Eq. 19.31.
546 19 Mixture of Ideal Gases

Thus, it is   
Vtotal pi = n i Mi R i T = R M T ni . (19.46)
i i i

In the mixture also applies 


pi = ptotal (19.47)
i

respectively 
n i = n total . (19.48)
i

A combination yields 
pi
i ptotal RM T
 = = . (19.49)
ni n total Vtotal
i

With
pi Vtotal
RM T = (19.50)
ni

it finally results in
pi ni
πi = = = xi . (19.51)
ptotal n total

In combination with Eq. 19.43 it is

pi Vi ni
= = = πi = σi = xi (19.52)
ptotal Vtotal n total

Hence, the partial pressure ratio πi is identical with the volume concentration σi as
well as with the molar concentration xi .

Gas Constant

In this section, a representative gas constant Rtotal is derived for a mixture of ideal
gases. For each gas within the mixture, see Fig. 19.3, it follows, cf. Eq. 19.42,

p1 Vtotal = m 1 R1 T
p2 Vtotal = m 2 R2 T
.. (19.53)
.
pn Vtotal = m n Rn T.
19.3 Laws of Mixing 547

Fig. 19.3 Gas constant of a


mixture of ideal gases

The summation of all components results in

( p1 + p2 + · · · + pn ) Vtotal = (m 1 R1 + m 2 R2 + · · · + m n Rn ) T. (19.54)

A representative gas constant is now defined as follows

!
(m 1 R1 + m 2 R2 + · · · + m n Rn ) T = m total Rtotal T. (19.55)

Thus, the thermal equation of state for the mixture reads as

ptotal Vtotal = m total Rtotal T (19.56)

Equation 19.55 can be applied to derive Rtotal , i.e.

m 1 R1 + m 2 R2 + · · · + m n Rn 
Rtotal = = ξi Ri (19.57)
m total i

Consequently, the representative gas constant of a mixture can be calculated with the
mass-averaged individual gas constants of its individual components.

Molar Mass

The molar mass of a mixture Mtotal shown in Fig. 19.3 is


m total  mi  ni
Mtotal = = = Mi . (19.58)
n total i
n total i
n total

Hence, by introducing the molar concentration xi , see Eq. 19.5, the molar mass of a
mixture finally is

Mtotal = (xi Mi ) (19.59)
i
548 19 Mixture of Ideal Gases

The representative molar mass of a mixture can thus be calculated using the molar
averaged individual molar masses of its single components. Both Rtotal and Mtotal
allow the application of the thermal equation of state for a mixture. The correlation
between individual gas constant Rtotal and general gas constant RM is

m total Rtotal = n total RM . (19.60)

Consequently, it finally is
RM = Rtotal Mtotal (19.61)

Conversion Mass/Molar Concentration

In this section, a correlation between the mass concentration ξi and the molar con-
centration xi is derived. Following the definition of the mass concentration
mi
ξi = (19.62)
m total

the mass can be replaced by molar quantity and molar mass, i.e.

n i Mi Mi
ξi = = xi . (19.63)
n total Mtotal Mtotal

Hence, it results in
Mi
ξi = xi (19.64)
Mtotal

19.3.2 Internal Energy, Enthalpy

Gibbs’ theorem is an extension of Dalton’s law, see [1]. It states that the total internal
energy of a mixture of ideal gases Utotal is equal to the sum of all partial internal
energies Ui of the components, i.e.
 
Utotal = U1 + U2 + · · · + Un = Ui = (m i u i ). (19.65)
i i

The specific internal energy of a mixture u total can thus be derived by dividing the
mass of the mixture, i.e.

Utotal 
u total = = (ξi u i ) (19.66)
m total i
19.3 Laws of Mixing 549

The enthalpy of a component i in a mixture of ideal gases is obtained according to


the definition of enthalpy, see Part I, i.e.

Hi = Ui + pi Vtotal . (19.67)

In a mixture, component i has a partial pressure pi and is distributed throughout the


volume Vtotal . The total enthalpy of a mixture is

Htotal = Utotal + ptotal Vtotal . (19.68)

According to Gibbs/Dalton, it is composed of its individual components, i.e.


   
Htotal = H1 + H2 + · · · + Hn = Hi = Ui + Vtotal pi = (m i h i ) .
i i i i
(19.69)
The specific enthalpy of a mixture h total can thus be derived by dividing the mass of
the mixture, i.e.
Htotal 
h total = = (ξi h i ) (19.70)
m total i

Specific Heat Capacities of a Mixture of Ideal Gases

For ideal gases the internal energy is exclusively a function of temperature, see
Chap. 12. Hence, it follows for a component i

dUi = m i cv dT. (19.71)

Thus, for a mixture it is



!
dUtotal = m i cv,i dT = m total cv,total dT. (19.72)
i

The isochoric specific heat capacity of a mixture cv,total is finally as follows

 mi 

cv,total = cv,i = ξi cv,i (19.73)


i
m total i

The enthalpy of a component i within a mixture of ideal gases obeys

dHi = m i c p dT. (19.74)


550 19 Mixture of Ideal Gases

Fig. 19.4 Adiabatic mixing closed system—exemplary for n = 3 components

Thus, for the mixture it is



!
dHtotal = m i c p,i dT = m total c p,total dT. (19.75)
i

Finally, the isobaric specific heat capacity of a mixture c p,total yields

 mi 

c p,total = c p,i = ξi c p,i (19.76)


i
m total i

19.3.3 Adiabatic Mixing Temperature

Closed System

For a closed system, see Fig. 19.4, the first law of thermodynamics obeys16

Q AB + WAB = U. (19.77)

Since the system is adiabatic and no work is exchanged, the first law of thermody-
namics simplifies to
U = 0. (19.78)

As shown in the previous section, the change in the internal energy of a mixture
is composed of the changes in the internal energy of its individual components as
follows17

16 Kinetic as well as potential energies are ignored.


17 Subject to the condition, that the specific heat capacity is temperature-independent.
19.3 Laws of Mixing 551

Fig. 19.5 Adiabatic mixing open system—exemplary for n = 3 components

U1 = m 1 cv,1 (TB − T1 )


U2 = m 2 cv,2 (TB − T2 )
.. (19.79)
.
Un = m n cv,n (TB − Tn ).

The total change in internal energy U is therefore the sum of its individual compo-
nents i, i.e.  
U = Ui = m i cv,i (TB − Ti ) = 0. (19.80)
i i

Rearranging results in
 
TB m i cv,i − m i cv,i Ti = 0. (19.81)
i i

This equation can be solved for the temperature TB after mixing in state (B), i.e.

m i cv,i Ti
i
TB =  (19.82)
m i cv,i
i

Open System

For an open system in steady state, see Fig. 19.5, the first law of thermodynamics
obeys18

Q̇ AB + PAB =  Ḣ . (19.83)

18 Kinetic as well as potential energies are ignored.


552 19 Mixture of Ideal Gases

Since the system is adiabatic and no power is exchanged, the first law of thermody-
namics simplifies to
 Ḣ = 0. (19.84)

As shown in the previous section, the enthalpy change of a mixture is composed of


the changes in the enthalpies of its individual components as follows19

 Ḣ1 = ṁ 1 c p,1 (TB − T1 )


 Ḣ2 = ṁ 2 c p,2 (TB − T2 )
.. (19.85)
.
 Ḣn = ṁ n c p,n (TB − Tn ).

The total change in enthalpy Ḣ is therefore the sum of its individual components i,
i.e.  
 Ḣ =  Ḣi = ṁ i c p,i (TB − Ti ) = 0. (19.86)
i i

Rearranging results in
 
TB ṁ i c p,i − ṁ i c p,i Ti = 0. (19.87)
i i

This equation can be solved for the temperature TB after mixing in state (B), i.e.

ṁ i c p,i Ti
i
TB =  (19.88)
ṁ i c p,i
i

19.3.4 Irreversibility of Mixing

Entropy

Although the irreversibility of mixing has already been derived on the basis of the
second law of thermodynamics in conjunction with the partial energy equation, see
Sect. 19.2, an alternative approach is now presented.
In order to analyse the irreversibility of an adiabatic, isothermal mixing of ideal
gases,20 cf. Fig. 19.6, the generation of entropy from state (A) to state (B), i.e.

19Subject to the condition, that the specific heat capacity is temperature-independent.


20In this case, two ideal gases are mixed for simplicity. However, this approach can easily be
extended to n components.
19.3 Laws of Mixing 553

Fig. 19.6 Irreversibility of mixing (irreversible, adiabatic)

Fig. 19.7 Irreversibility of mixing—equivalent model (reversible, diabatic)

Fig. 19.8 Thought experiment for a reversible, isothermal mixing: pistons replace separating wall
in Fig. 19.2

Si,12 , is required. An equivalent model is now postulated that allows for hypothetical
reversible mixing.21 However, even with this hypothetical mixing, the same state (B)
is to be achieved as with the actual irreversible mixing, see state (B) in Fig. 19.6. A
technique for this thought experiment is shown with both Fig. 19.7 as well as with
Fig. 19.8. Since states (A) and (B) have the same temperature, the equivalent model
investigates an isothermal mixing, see Fig. 19.7. Instead of separating the ideal gases

21 This is just hypothetic, since mixing of different gases is always irreversible.


554 19 Mixture of Ideal Gases

by a wall, the gases are now in two chambers bounded by two pistons. Each of
the pistons can move freely, i.e. without any friction. As the pistons move slowly
towards the cylinder walls, mixing occurs due to diffusion of the gases, as shown in
Fig. 19.8: Gas 1 and gas 2 expand into the gap and further displace pistons 1 and 2.
Piston 1 is therefore permeable to gas 1 and impermeable to gas 2. Piston 2 behaves
vice versa, i.e. gas 2 can pass through the piston and gas 1 cannot. This selective
behaviour of the pistons is the key parameter of reversible mixing in this equivalent
model. Obviously, the two gases expand22 during this change of state so that the gases
release work to the moving pistons: The left side of piston 1 allows gas 1 to pass,
so that gas 1 does not exert a force on piston 1. In the filling chamber, however, it is
gas 2 that exerts a force on piston 1, since gas 2 cannot pass piston 1. Consequently,
work must be released due to the effective force and the distance that piston 1 moves.
The force on piston 2 follows accordingly and is caused by gas 1. In sum, the volume
work of gas 1 is transferred by piston 2, the volume work of gas 2 by piston 1.
If no irreversibilities23 occur during the mixing in the cylinder, the amount of work
required for this expansion equals exactly the amount of work required conversely for
pushing in the pistons, i.e. compressing the two gases.24 During such compression,
gases 1 and 2 are separated again due to the characteristics of the pistons. The entire
process is therefore reversible, since the state (A) has been reached again and the
environment does not change. Let us now concentrate on the change of state from
(A) to (B): The total work released in the process is the isothermal25 volume work
of the two gases and is as follows26

Vtotal Vtotal
Wtotal = WV,1 + WV,2 = − ptotal V1 ln − ptotal V2 ln < 0, (19.89)
V1 V2

since the change of state shall be isothermal, see also Sect. 19.2. To keep the change
of state isothermal, according to the first law of thermodynamics, heat needs to be
supplied, i.e.
Wtotal + Q = U2 − U1 = 0. (19.90)

Obviously, in order to achieve state (B), the heat supply needs to follow

Q = −Wtotal > 0. (19.91)

With the supplied heat, the system receives additional entropy. Thus, the reversible
mixing follows

22 Expansion means release of volume work.


23 This requires a quasi-static change of state.
24 According to the partial energy equation.
25 The equivalent model investigates an isothermal mixing.
26 See Fig. 13.13.
19.3 Laws of Mixing 555

Q Wtotal ptotal Vtotal V1 Vtotal V2 Vtotal
Sa = =− = ln + ln .
T T T Vtotal V1 Vtotal V2
(19.92)
Hence, the entropy in state (B) obeys

SB = SA + Sa . (19.93)

The premise is that the irreversible mixing and the equivalent model of the hypothet-
ical reversible mixing reach the same state (B). Thus, both cases must reach the same
entropy increase. In case of irreversible mixing, there has been no heat exchange and
thus no entropy transport Sa . In order to achieve the same amount of entropy, entropy
must therefore be generated, i.e.

SB = SA + Si . (19.94)

A comparison of Eqs. 19.93 and 19.94 leads to Si = Sa , so that



ptotal Vtotal V1 Vtotal V2 Vtotal
Si = ln + ln . (19.95)
T Vtotal V1 Vtotal V2

Applying the thermal equation of state and Dalton’s law, see Eqs. 19.21 and 19.24,
the generation of entropy for an isothermal mixing of two ideal gases results in

ptotal ptotal
Si = m 1 R1 ln + m 2 R2 ln (19.96)
p1 p2

Thus, the same result is achieved as with Eq. 19.39. Note the so-called Gibbs paradox,
which states that no mixing entropy occurs when two identical ideal gases are mixed,
i.e. Si,Mix = 0. In case the two compartments contain the same ideal gas at pressure
p and temperature T , there is a mixing when the sliding wall between the two
compartments opens. When closing this wall again, each compartment then possesses
the same state as before. Hence, from this perspective the change of state needs to be
reversible. However, according to conventional thermodynamics there needs to be
generation of entropy while the sliding wall opens and the two identical ideal gases
mix with each other. Since the initial state is obviously achieved, when closing the
wall, the amount of entropy needs to be the same as before. Thus, entropy must have
been destroyed when closing the wall again,27 i.e. the second law of thermodynamics
is violated. In order to solve this problem a correction term needs to be defined
that takes into account the permutation of identical particles by means of quantum
mechanics. For further explanation see [36].
Based on the caloric equation of state in terms of entropy,28 which has already
been introduced in Part I,

27 Otherwise, it is not possible to get the same amount of initial entropy.


28 Note that s0 is the specific entropy at an arbitrary reference level T0 , p0 .
556 19 Mixture of Ideal Gases

T1 p1
s1 − s0 = c p ln − R ln (19.97)
T0 p0

the entropy of state (A) follows accordingly:


• Entropy of component 1:

T ptotal
SA,1 = m 1 s0 + m 1 c p,1 ln − m 1 R1 ln (19.98)
T0 p0

• Entropy of component 2:

T ptotal
SA,2 = m 2 s0 + m 2 c p,2 ln − m 2 R2 ln (19.99)
T0 p0

Hence, the total entropy in state (A) is

SA = SA,1 + SA,2 . (19.100)

Entropy of a Mixture

The entropy after adiabatic mixing in state (B) therefore results in

SB = SA + Si . (19.101)

A combination of Eq. 19.101 with Eqs. 19.100, 19.99, 19.98 and 19.96 leads to

T ptotal
SB =m 1 s0 + m 1 c p,1 ln − m 1 R1 ln
T0 p0

T ptotal
+ m 2 s0 + m 2 c p,2 ln − m 2 R2 ln (19.102)
T0 p0

ptotal ptotal
+ m 1 R1 ln + m 2 R2 ln .
p1 p2

Thus, by rearranging

T p1
SB = m 1 s0 + m 1 c p,1 ln − m 1 R1 ln
T0 p0
  
S1
(19.103)
T p2
+ m 2 s0 + m 2 c p,2 ln − m 2 R2 ln .
T0 p0
  
S2
19.3 Laws of Mixing 557

In general, the entropy of a mixture of n ideal gases in thermodynamic equilibrium,


i.e. in state (B), is the sum of its partial entropies Si , which are calculated with the
respective partial pressures pi and the equilibrium temperature T :


n
SB = Si ( pi , T ) (19.104)
i

If a change in entropy is calculated, e.g. if a gas mixture changes its state, all reference
values, i.e. s0 , T0 and p0 , disappear.

Exergy

In this section, it is investigated how much exergy the isothermal, isobaric mixing,
see Fig. 19.6, possesses. Since the pure mixing is in focus, the system shall have
already reached ambient temperature Tenv , i.e. it shall be in thermal equilibrium. The
heat transfer to bring the system from temperature T to ambient temperature Tenv is
not of interest. Since the temperature is constant and the masses of the gases in the
volumes V1 and V2 are fixed, the pressure of the system is also fixed. Hence, once the
pressure is adjusted, e.g. by the amount of gases 1 and 2, it is not possible to bring
the system to ambient pressure under the given premises. A hypothetical reversible
mixing would release volume work as the two gases expand from state (A) to state
(B). Equation 19.89 gives the maximum work that can hypothetically be released
by the system if the change of state were reversible, i.e. this work is identical to the
exergy29 of the mixture:

Vtotal Vtotal
E x,mix = −Wtotal = + ptotal V1 ln + ptotal V2 ln >0 (19.105)
V1 V2

Rearranging and applying the thermal equation of state results in



Vtotal Vtotal
E x,mix = m 1 R1 Tenv ln + m 2 R2 Tenv ln . (19.106)
V1 V2

Application of Dalton’s law, see Eqs. 19.21 and 19.24, finally leads to

ptotal ptotal
E x,mix = m 1 R1 Tenv ln + m 2 R2 Tenv ln (19.107)
p1 p2

The mixing according to Fig. 19.6, i.e. irreversible mixing, has shown, that dissipation
occurs, see Sect. 19.2. The dissipation can be calculated with Eq. 19.33, so that the

29 Note that the effective work, see Sect. 9.2.3, does not play a role, since the pistons are encapsulated
from the environment—except for the negligible cross sections of the two rods.
558 19 Mixture of Ideal Gases

entropy generation has been calculated as follows


ptotal ptotal
Si = m 1 R1 ln + m 2 R2 ln . (19.108)
p1 p2

Hence, the loss of exergy due to the mixing is

ptotal ptotal
E x,V = Tenv Si = m 1 R1 Tenv ln + m 2 R2 Tenv ln (19.109)
p1 p2

A comparison of Eqs. 19.109 and 19.107 shows, that the entire exergy gets lost during
the mixing process, i.e.
E x,V = E x,mix . (19.110)

For a mixture of n ideal gases, the following applies accordingly


n
ptotal
E x,V = E x,mix = m i Ri Tenv ln (19.111)
i
pi

Problem 19.1 A tank with a volume V = 1 m3 initially, i.e. state (A), contains air
with pA = 1 bar and 20 ◦ C. Now hydrogen (H2 ) is filled in under pressure so that the
tank finally reaches state (B) with pB = 4 bar and 20 ◦ C, see Fig. 19.9.
(a) What are the volume fractions of air and hydrogen in the tank?
(b) Determine the gas constant and molar mass of the mixture.
(c) How many moles of the mixture are in the tank?
Further information:
kg kg kg
• Molar masses: MH2 = 2 kmol , MO2 = 32 kmol , MN2 = 28 kmol
• General gas constant: RM = 8.3143 kmol K
kJ

• Composition of air: Nitrogen 79 vol. − %, oxygen 21 vol. − %

Solution

(a) Volume fractions


• Air (1)

In state (A)30 it is
pA V = m 1 R1 T. (19.112)

In state (B) it is

30 In state (A), the total pressure is caused exclusively by air.


19.3 Laws of Mixing 559

Fig. 19.9 Sketch to Problem 19.1

pB,1 V = m 1 R1 T. (19.113)

Thus, it follows
pB,1 = pA = 1 bar. (19.114)

The partial pressure follows accordingly and is equal to the volume ratio

pB,1
σB,1 = πB,1 = = 0.25 (19.115)
pB

• Hydrogen (2)

The total pressure is the summation of the partial pressures in state (B), i.e.

pB,1 + pB,2 = pB . (19.116)

Hence, the partial pressure of the hydrogen is

pB,2 = 3 bar. (19.117)

The partial pressure follows accordingly and is equal to the volume ratio

pB,2
σB,2 = πB,2 = = 0.75 (19.118)
pB

(b) Gas constant and molar mass of the mixture

The molar mass for hydrogen (2) is

kg
M H2 = M 2 = 2 . (19.119)
kmol
560 19 Mixture of Ideal Gases

Its gas constant is


kJ
RM 8.3143 kmol kJ
RH2 = R2 = = kg
K
= 4.1571 . (19.120)
M2 2 kJ kg K

For air (1) with a composition of σ1,O2 = 0.21, σ1,N2 = 0.79 the molar mass
follows
  kg kg
M1 = x i Mi = σ Mi = (0.21 · 32 + 0.79 · 28) = 28.84 .
i i
kmol kmol
(19.121)
The gas constant for air (1) is
RM kJ
R1 = = 0.28829 . (19.122)
M1 kg K

Now that the components Ri and Mi respectively are known, the properties of
the mixture can be determined. For the molar mass of the mixture MB it follows
  
MB = x i Mi = σi Mi = πi Mi . (19.123)
i i i

Thus, for state (B)

kg
MB = πB,1 M1 + πB,2 M2 = 8.71 (19.124)
kmol

The gas constant for the mixture RB reads as

RM J
RB = = 954.56 (19.125)
MB kg K

(c) Molar amount of the mixture

The thermal equation of state can be applied for state (B), i.e.

pB V = m B R B T (19.126)

respectively
pB V = n B MB RB T = n B RM T. (19.127)

Thus, the molar quantity is

pB V
nB = = 0.1641 kmol (19.128)
RM T
19.3 Laws of Mixing 561

Problem 19.2 A mixture of 32 mass−% ammonia (RNH3 = 0.4882 kgkJK , κNH3 =


1.31) and 68 mass−% nitrogen (RN2 = 0.2968 kgkJK , κN2 = 1.40) with a temperature
of 10 ◦ C and a total mass of 6.3 kg is compressed reversibly and adiabatically from
0.98 bar to 2.94 bar.
(a) What is the molar mass of the mixture?
(b) Calculate the specific isochoric heat capacity cv of the mixture!
(c) What is the change of enthalpy that is related with the compression?
General gas constant: RM = 8.3143 kmol
kJ
K
.

Solution

(a) The molar mass of the mixture follows

RM
M= . (19.129)
R
The gas constant of the mixture R can be easily calculated, since the mass
concentrations are given
 kJ kJ
R= ξi Ri = (0.32 · 0.4882 + 0.68 · 0.2969) = 0.358048 .
i
kg K kg K
(19.130)
Thus, the molar mass is

RM kg
M= = 23.22 (19.131)
R kmol

(b) The specific isochoric heat capacity cv of the mixture can be calculated according
to 
cv = ξi cv,i . (19.132)
i

For the single components it is known from Part I, that

Ri
cv,i = . (19.133)
κi − 1

Thus, for ammonia and nitrogen it is

RNH3 kJ
cv,NH3 = = 1.5748 (19.134)
κNH3 − 1 kg K

respectively
562 19 Mixture of Ideal Gases

RN2 kJ
cv,N2 = = 0.742 . (19.135)
κN2 − 1 kg K

Hence, the specific isochoric heat capacity cv of the mixture follows

 kJ
cv = ξi cv,i = ξNH3 cv,NH3 + ξN2 cv,N2 = 1.0085 (19.136)
i
kg K

(c) To calculate the enthalpy change associated with compression, the temperature
change must first be determined. Since the change of state is adiabatic and
reversible, it is isentropic. An isentropic change of state31 is, see Part I,
κ−1
κ
p2
T2 = T1 . (19.137)
p1

Thus, the isentropic coefficient of the mixture κ needs to be calculated. For the
mixture it is
kJ
c p = R + cv = 1.366548 . (19.138)
kg K

Hence, the isentropic coefficient of the mixture follows


cp
κ= = 1.35503. (19.139)
cv

The temperature after the compression is


κ−1
κ
p2
T2 = T1 = 377.59 K. (19.140)
p1

The temperature rise is

T = T2 − T1 = 94.445 K. (19.141)

For the change of specific enthalpy the caloric equation of state32 can be applied,
i.e.
dh = c p dT. (19.142)

Its integration yields33

31 This is true because the mixture consists of ideal gases.


32 Since the mixture consists of ideal gases, the total specific enthalpy is exclusively a function of
temperature.
33 Since no temperature dependencies are given.
19.3 Laws of Mixing 563

kJ
h = c p T = 129.06 . (19.143)
kg

A multiplication with the total mass leads to the entire change of enthalpy, i.e.

H = m h = 813.1 kJ (19.144)
Chapter 20
Humid Air

The previous chapter has clarified how to treat mixtures of ideal gases. Definitions
for characterising the mixture composition have been introduced as well as the irre-
versibility that arises when mixing different ideal gases. In this chapter, however,
the focus is on a technically relevant mixture of fluids: humid air. Its components at
atmospheric conditions are as follows:
• Dry air, i.e. a mixture of N2 and O2 and with a small amount of CO2 . This type
of mixture has been treated in Chap. 19. Mostly, these components are treated as
ideal gases, so that the principles of Chap. 19 can be easily applied.
• Water can occur in humid air as a liquid, i.e. as mist or rain, as a solid, i.e. as ice or
snow, and as vapour. Chapter 18 has shown, that water as a real fluid can be subject
to a phase change. Our everyday experience shows that this phase change takes
place under atmospheric conditions such as rain or snow. It is therefore necessary
to clarify how the mixture can be described when vapour condenses and forms a
liquid.
Both, dry air and vapour, form a gaseous phase1 and are commonly treated as ideal
gases,2 see Fig. 20.1a. According to Dalton, the total pressure of humid air is the
sum of the partial pressures of dry air pa and vapour pv , i.e.

p = pa + pv . (20.1)

The deviation from non-ideal gas behaviour is discussed in this chapter. However,
this assumption leads to an error less than 3% for temperatures below 100 ◦ C. Usually
this error is acceptable for HVAC3 systems.

1 Usually, the partial pressure of vapour is significantly smaller than the total pressure, i.e. pv  p.
2 This figure shows air in a so-called unsaturated state.
3 Heating, Ventilation and Air Conditioning.

© Springer Nature Switzerland AG 2022 565


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_20
566 20 Humid Air

Fig. 20.1 Humid air—a unsaturated, b saturated, c saturated with additional liquid/ice

If the water content in the air continues to increase, the gas phase reaches a so-
called saturated state in which the vapour content becomes maximum, see Fig. 20.1b.
If additional water is supplied to the system, it occurs in liquid or solid state depending
on the temperature. In thermodynamic equilibrium, the gas phase, consisting of dry
air and vapour, is still in a saturated state, see Fig. 20.1c.
Therefore, this chapter examines, among others, the boundary conditions under
which a phase change takes place and how it influences the composition of the gas
mixture of dry air and vapour and the energy balance.
Finally, it is discussed how and with which technical aids air can be conditioned
in terms of temperature and humidity and what energy supply is associated with
conditioning.

20.1 Thermodynamic State

20.1.1 Concentration

The water content of air4 x is defined as the ratio of the mass of water in the air to
the mass of dry air, i.e.
mw
x= . (20.2)
ma

Not to be confused with the molar concentration x introduced in Chap. 19 or the


vapour ratio x defined in Chap. 18.

Since the mass of water m w in the air can occur as vapour (v), liquid (liq) or ice
(i), it is
m w = m v + m liq + m ice . (20.3)

Thus, the water content follows

4 Sometimes also denoted as moisture grade.


20.1 Thermodynamic State 567

mw mv m liq m ice
x= = + + = xv + xliq + xice (20.4)
ma ma ma ma

Thermodynamic states in which water occurs simultaneously as vapour, liquid and


solid are possible and are discussed later. The following examples show that the
water content x can be any positive number:
• Dry air
mw 0
x= = =0 (20.5)
ma ma

• Pure liquid water


mw m liq
x= = →∞ (20.6)
ma 0

20.1.2 Aggregate State of the Water

To decide on the state of aggregation of water in air, both its pressure and its tempera-
ture are required. Chapter 18 has shown that these two state values determine the state
of aggregation. The state of water can easily be visualised e.g. in a p, T -diagram,5
cf. Fig. 20.2, and is further explained: Such a diagram shows that water can occur
as vapour or steam, as a liquid or as a solid, depending on the pressure and tem-
perature. The vapour pressure curve ps = ps (T ) marks the boundary line between
vapour/steam and liquid and is important for the characterisation of humid air. Obvi-
ously, vaporisation at low pressures already starts at low temperatures. However, the
wet steam region, cf. Sect. 18.2, cannot be seen in this p, T -projection because it is
covered by the vapour pressure line, cf. Sect. 18.2.2. In this two-phase region, the
pressure is a function of the temperature, i.e. further state information is required,
e.g. the vapour ratio, to unambiguously determine the thermodynamic state.
If dry air comes into contact with a water reservoir, e.g. in a liquid state, a ther-
modynamic imbalance6 occurs, i.e. a high concentration of water on the one side
and no water on the other side. According to the second law of thermodynamics,
this concentration gradient causes evaporisation7 of the water in order to balance
the concentration gradient: Some water molecules can escape from the water sur-
face due to their kinetic energy, i.e. their temperature. These water molecules are
present in the air as a vapour phase due to a lower binding than in the liquid, see also
Sect. 20.1.3. Apparently, the partial pressure of water in the surrounding air is small.

5 Note that the p, T -diagram only takes into account the water, the air is not part of this diagram.
6 This imbalance is based on a chemical potential, see Example 24.5.
7 It needs to be distinguished between vaporisation and evaporation. Vaporisation means the phase

change of a single fluid, e.g. water, that was handled in Chap. 18. Evaporation, on the other hand,
describes the vaporisation of one fluid, e.g. water, and its diffusive transport into a second gas, e.g.
air.
568 20 Humid Air

Fig. 20.2 p, T -diagram of


water

The transport mechanism of the water molecules at the liquid/air interface is based
on diffusion and has already been explained in the previous chapter.
Obviously, the phase change of the water takes place directly at this interface:
According to Fig. 20.2, water with a low partial pressure can be in vapour state
depending on its temperature, while the water reservoir is in liquid state. The more
water molecules leave the reservoir, the higher the partial pressure of the vapour in
the ambient air becomes. This continues until the saturated state is reached accord-
ing to the vapour pressure curve. However, the required energy for the evapora-
tion/vaporisation is taken from the environment and reduces its temperature.8 The
energy balance of such a change of aggregation state is discussed later in this chapter.
The more water vaporises, the higher the partial pressure of the water and the lower
the partial pressure of the air, i.e. the vapour concentration increases. Nevertheless,
according to Dalton, the following applies to the total pressure of the gas phase9

p = pa + pv . (20.7)

Let us assume, that air and water both have the same temperature T1 . Following the
vapour pressure curve ps = ps (Ts ) in Fig. 20.2, the maximum vapour pressure is
limited, i.e. the vaporisation at temperature T1 stops as soon as the partial pressure
of the vapour reaches ps (T1 ). The supply of further water into the humid air means
that this water cannot be carried as vapour, i.e. condensation leads to liquid water or
even ice.

The following classification applies to air with an exemplary temperature T1 :


• Unsaturated humid air:
In the case
pv < ps (T1 ) (20.8)

8 Also known as evaporative cooling.


9 In thermodynamic equilibrium, the gas phase applies its pressure p on the liquid.
20.1 Thermodynamic State 569

the air is still capable of carrying a higher amount of vapour. It is therefore unsat-
urated.
• Saturated humid air:
In the case
pv = ps (T1 ) (20.9)

the maximal load of vapour is reached. Any further supply of water cannot be
carried as vapour, but occurs as liquid or solid water. But even if further water is
supplied, the gas phase in thermodynamic equilibrium still contains the maximum
load of vapour belonging to ps (T1 ).

20.1.3 Distinction Between Vaporisation and Evaporation

With the following illustration, see Fig. 20.3, the process of phase change is examined
in more detail. In the first chapters of Part II the phase change for a single-component
fluid has been investigated. This process takes place, for example, at the interface
between a hot plate and a fluid above it. By supplying thermal energy, it is possible
to overcome the binding forces between the fluid particles, so that the fluid is finally
perceived as a vapour bubble. However, the single-component system in case (a) still
consists only of the chemical compound undergoing the phase change, for example
water, i.e. H2 O.
At the upper interface, the transition between the first component, e.g. water in
liquid state, and the surrounding air as the second component, is now to be considered,
see Fig. 20.3b. Within the liquid, the particles experience high binding forces among

Fig. 20.3 Distinction between a vaporisation and b evaporation


570 20 Humid Air

themselves, so that the water is perceived as liquid. At the interface with the air, the
binding forces become lower because there is no other upper layer of liquid water: At
molecular level, the particles of a fluid are in motion and follow a velocity distribution,
i.e. they possess specific kinetic energy that increases with increasing temperature.
Particles which have a high kinetic energy can overcome the binding forces at the
water/air phase interface, see Fig. 20.3b. Vice versa, water molecules that are in the
gas phase with the air and have a low specific kinetic energy are recaptured by the
liquid. Molecules being able to leave the liquid are then present as vapour.10 Above
the phase boundary, there is now a gas phase consisting of air and vapour. Due to
the imbalance with regard to the water concentration (a lot of water below the phase
boundary, little water above), a balancing process now occurs, i.e. water particles
pass into the air phase as vapour until a thermodynamic equilibrium is reached: The
partial pressure of the water at the phase boundary reaches the saturation pressure that
belongs to the present equilibrium temperature. The process described in Fig. 20.3b
is also known as evaporation.
Vaporisation in a liquid thus describes the phase change liquid/gas depending on
the pressure under which the liquid is present, e.g. at 1bar at approx. 100 ◦ C for
water. Evaporation, in contrast, describes the phase change in the presence of an
additional gas phase, e.g. air. Here, the partial pressure of the water in the gas phase
is decisive for the thermodynamic state at which the phase change takes place, i.e.
the evaporation process can already start at low temperatures, cf. Tables A.1, A.2,
A.3 and A.4.

20.1.4 Unsaturated Versus Saturated Air

Vapour content and partial pressure

To find a correlation between partial pressure and vapour content, it is assumed


that both gases, dry air and vapour, can be considered ideal. Applying the thermal
equation of state for ideal gases for dry air leads to

pa V
pa V = m a R a T → m a = (20.10)
Ra

and vapour accordingly

pv V
pv V = m v R v T → m v = . (20.11)
Rv

The vapour content of the air is by definition, see Sect. 20.1.1,

10 A liquid is characterised by strong binding forces, while gaseous state means low binding forces.
20.1 Thermodynamic State 571

mv
xv = . (20.12)
ma

A combination of Eqs. 20.10 and 20.11 with Eq. 20.12 results in

pv R a
xv = . (20.13)
pa R v

The partial pressure of air pa follows from Dalton’s law, see Eq. 20.7, and can be
written as
pa = p − pv . (20.14)

Substituting Eq. 20.13 into Eq. 20.14 yields

pv R a
xv = . (20.15)
( p − pv ) R v

With the individual gas constants for dry air11

RM J
Ra = = 287.0409 (20.16)
Ma kg K

with
kg
Ma = xO2 MO2 + xN2 MN2 + xAr MAr + xCO2 MCO2 = 28.9656 (20.17)
kmol

and vapour12
RM J
Rv = = 461.5158 (20.18)
MW kg K

results in
pv
xv = 0.622 (20.19)
p − pv

Respectively, converted to the partial pressure of the vapour as a function of the water
content
xp
pv = (20.20)
0.622 + xv

kg kg
11 The molar masses of the components of air are MO2 = 31.998 kmol , MN2 = 28.0134 kmol , MAr =
kg kg
39.948 kmol and MCO2 = 44.0087 kmol . The molar fractions follow according to Chap. 19.
12 With the molar mass for water M = 18.0152 kg .
W kmol
572 20 Humid Air

Maximum vapour load of humid air

If the humid air is saturated, the partial pressure of the vapour is equal to the saturation
vapour pressure, see Fig. 20.2. All variables that are in the saturated state are from
now on marked with a dash ( ). Thus, it is

pv = ps (T ) . (20.21)

Hence, following Eq. 20.19 the maximum vapour content of humid air is

pv
xv = 0.622 (20.22)
p − pv

According to this equation, the maximum vapour content is a function of the total
pressure p and the temperature T of the humid air, i.e. the greater the temperature,
the greater the maximum vapour load.13
Vice versa, one gets the partial pressure in the saturated state from the maximum
vapour content xv , i.e.
xv · p
pv = (20.23)
0.622 + xv

Relative saturation/humidity

If humid air is unsaturated, it contains less vapour than is potentially possible at


its current temperature T . The ratio of the current vapour content to its maximum
content at its temperature T is called the degree of saturation or relative saturation
ψ, i.e.
xv
ψ=  (20.24)
xv

Another important measure of humid air is the relative humidity ϕ, which is defined
as pv
ϕ= . (20.25)
pv

Applying Eq. 20.20 finally results in

pv xv p
ϕ= = · (20.26)
pv xv + 0.622 pv

The correlation between degree of saturation ψ and the relative humidity ϕ follows
by applying Eqs. 20.19 and 20.22, i.e.

13 This is due to the saturation vapour pressure increasing with rising temperature, see Fig. 20.2.
20.1 Thermodynamic State 573

xv pv p − pv pv p − pv
ψ= 
= · 
=  · . (20.27)
xv p − pv pv pv p − pv

Thus, it finally is
ψ p − pv
= (20.28)
ϕ p − pv

Absolute humidity

The absolute humidity ρv is defined as the ratio of the mass of vapour m v to the total
volume of humid air V shared by vapour and dry air, i.e.
mv
ρv = . (20.29)
V
Applying the thermal equation of state results in

pv V 1 pv
ρv = · = (20.30)
Rv T V Rv T

One obtains the maximum absolute humidity ρv when the air is saturated, so that
pv = ps (T )
p
ρv = v . (20.31)
Rv T

Hence, the maximum mass of the vapour within the saturated air follows

pv
m v = · V. (20.32)
Rv T

Example 20.1 Starting from the initial state (1), characterised by pa,1 = 983 mbar
and ϑ1 = 20 ◦ C, humid air is cooled at constant total pressure, i.e. p = 1 bar, and
constant water content.

What does the change of state look like?

According to Dalton, the partial pressure of the vapour is obtained as follows

pv,1 = p − pa,1 = 17 mbar. (20.33)

The maximum partial pressure of the vapour at ϑ1 = 20 ◦ C can be taken from the
steam table, cf. Table A.1, and amounts to

pv,1 = ps (ϑ1 ) = 23.392 mbar. (20.34)
574 20 Humid Air

Fig. 20.4 Humid


air—Example 1/2

The humid air in state (1) is therefore unsaturated, since it lies below the vapour
pressure curve in the p, T -diagram and

pv,1 < pv,1 . (20.35)

However, when cooling with constant total pressure, the ability of the air to carry
vapour decreases, see the course of the ps = ps (T ) function. The partial pressures
of air and vapour are constant, since air and vapour content do not change, i.e.
pv nv
πv = = xv = = const. (20.36)
p na + nv

with p = const. Consequently, the change of state in Fig. 20.4 follows a horizontal
line as the temperature decreases and the partial pressure of the vapour14 remains
constant, i.e.
pv = pv,1 . (20.37)

This continues until state (2) is reached, which still has the same partial pressure of
the vapour as before, but is in saturated state, i.e.

pv,2 = pv,1 = 17 mbar = ps (T2 ) . (20.38)

According to temperature T2 , the air now carries the maximum load of vapour. Equa-
tion 20.38 can be solved using the steam table and ϑ2 = 14.95 ◦ C is obtained.

Starting from state (2) isobaric cooling continues. What happens?

14Only the vapour pressure is part of the p, T diagram in Fig. 20.4, although the y-axis is labelled
p. Hence, p does not represent the total pressure in this figure.
20.1 Thermodynamic State 575

Fig. 20.5 Humid air—Example 2/2

The change of state15 is shown in Fig. 20.5. State (1) has been unsaturated, while
state (2) is saturated. As the temperature decreases during cooling, the ability to carry
vapour also decreases. The further the temperature drops, the more water begins to
condense. During this process, the gas phase contains a smaller amount of vapour
than before, while the mass of dry air remains unchanged. Therefore, the partial
pressure of the vapour decreases while the partial pressure of the dry air increases.16
The total volume must decrease. This can be shown with the thermal equation of
state for dry air, i.e.
m a Ra T
V = . (20.39)
pa

The mass of the dry air m a remains constant, as does the gas constant Ra , while the
temperature T decreases and the partial pressure of the air pa increases. Therefore,
the volume V occupied by both dry air and steam must decrease. Since an increasing
reservoir of liquid water or ice is available during cooling, the humid air is always
in a saturated state,17 i.e.

pv,3 = pv,3 = ps (ϑ3 ) (20.40)

respectively, when water starts to freeze



pv,4 = pv,4 = ps (ϑ4 ) . (20.41)

Due to the abundant supply of water (solid/liquid), the water evaporates into the
gas phase until the air is saturated in thermodynamic equilibrium, i.e. the balancing
process eventually comes to a halt and saturated air and liquid or icy water are present.

15 The total pressure remains constant.


16 Otherwise, it would not be possible to keep the total pressure constant.
17 The gas phase is in saturated state, with an extra amount of liquid water or ice, see Fig. 20.5.
576 20 Humid Air

20.2 Specific State Values

It is already known from Part I that extensive state variables Z can be referred to the
mass of the system m in order to obtain mass-specific state variables z, i.e.

Z
z= . (20.42)
m
Part I dealt with simplified fluid models such as ideal gases or incompressible fluids.
However, the definition for specific state variables is independent of the type of fluid,
i.e. the definition can also be applied to real fluids or mixtures. It is now important
to note that in Eq. 20.42 m represents the associated total mass. However, when
calculating specific state values for humid air, a different approach is taken. By
definition, not the total mass of the gas mixture, but the mass of the dry air air m a is
used to calculate the mass-specific state variables. This is marked with the index18
1 + x, i.e.
Z
z 1+x = (20.43)
ma

This leads to the


• specific volume
V
v1+x = (20.44)
ma

• specific internal energy


U
u 1+x = (20.45)
ma

• specific enthalpy
H
h 1+x = (20.46)
ma

• specific entropy s1+x


S
s1+x = (20.47)
ma

for humid air.

18 Note that Z is the total extensive state value, i.e. of dry air and vapour.
20.2 Specific State Values 577

20.2.1 Thermal State Values

Specific volume of humid air

The specific volume of humid air with x ≤ xv is

V
v1+x = . (20.48)
ma

The partial volume of air Va and of vapour Vv form the total volume19 V of the
gaseous phase of humid air. Thus, it follows

Va Vv
v1+x = + . (20.49)
ma ma

By applying of the ideal gas law, i.e. pV = m RT , the following expressions can be
substituted in Eq. 20.49
Va Ra T
= (20.50)
ma p

and
Rv T
Vv = m v . (20.51)
p

Hence, the specific volume of unsaturated humid air results in


 
Ra T Rv T Ra T Rv
v1+x = + xv · = 1+ · xv . (20.52)
p p p Ra

The gas constants for dry air Ra and vapour Rv are constant and have been introduced
in Sect. 20.1.4. Thus, Eq. 20.52 simplifies to

Ra T
v1+x = [1 + 1.608 · xv ] (20.53)
p

According to the definition of Part I, i.e. with the total mass of the system, the specific
volume is
V
v= . (20.54)
m
Conversion leads to

19 According to Dalton, both gas phases, i.e. dry air and vapour, occupy the entire volume. However,
the partial volumes Va and Vv result if the gases are separated according to Fig. 19.2.
578 20 Humid Air

V V ma V ma 1
v= = · = · = v1+x · . (20.55)
ma + mv ma + mv ma ma ma + mv 1 + xv

Thus, the correlation between v and v1+x is

v1+x = (1 + xv ) · v (20.56)

The correlations above can also be applied if the saturated humid air contains small
amounts of liquid water or ice, respectively. This is due to the fact that the specific
volume of liquid/ice can be neglected compared to the volume of the gaseous phase,
cf. Fig. 20.1c.

Density of humid air

The density of humid air with x ≤ xv yields

1
ρ1+x = . (20.57)
v1+x

Thus, it is
 
p 1
ρ1+x = · (20.58)
Ra T 1 + 1.608 · xv

From
v1+x = (1 + xv ) · v (20.59)

thus follows the correlation20 between ρ1+x and ρ:

1 (1 + xv )
= (20.60)
ρ1+x ρ

with
ma + mv
ρ= . (20.61)
V

Example 20.2 Why does humid air rise in an atmosphere of dry air? The density of
humid air is, see Eq. 20.60,
 
p 1 + xv
ρ = (1 + xv ) · ρ1+x = · . (20.62)
Ra T 1 + 1.608 · xv

Two cases can therefore be examined, i.e.

20 Note that ρ considers the total mass.


20.2 Specific State Values 579

• density of dry air (xv = 0):


p
ρ (xv = 0) = (20.63)
Ra T

• density of humid air (xv > 0):


 
p 1 + xv
ρ (xv > 0) = · (20.64)
Ra T 1 + 1.608 · xv
     
ρ(xv =0) <1

Obviously, the density of humid air is smaller than the density of dry air

ρ (xv > 0) < ρ (xv = 0) . (20.65)

Thus, humid air rises in a dry atmosphere.

Problem 20.1 In a closed room a temperature of ϑ = 18 ◦ C and a relative humidity


of ϕ = 80% are measured. The total pressure is approximately standard pressure,
i.e. p = 1.01325 bar. The gas constant of dry air is Ra = 0.28704 kgkJK .

(a) Calculate the mass m of V = 1 m3 of the air.


(b) What is the maximum mass of vapour m v that the air with a volume V = 1 m3
and a temperature ϑ = 18 ◦ C can contain without forming mist?
Solution
(a) Total mass of the humid air m
• Approach 1:
Since the relative humidity is given, the partial pressure of the vapour pv can
be calculated. According to the steam table, cf. Table A.1 it is

ps (18 ◦ C) = 20.6466 mbar. (20.66)

Thus, the partial pressure of the vapour is

pv = ϕ · ps (18 ◦ C) = 16.5173 mbar. (20.67)

The vapour content xv follows


pv
xv = 0.622 = 0.010307 = 10.307 kgvapour /kgdry air . (20.68)
p − pv

This leads to
 
p 1 kg
ρ1+x = · = 1.1927 3 (20.69)
Ra T 1 + 1.608 · xv m
580 20 Humid Air

and hence to
kg
ρ = ρ1+x (1 + xv ) = 1.2050 . (20.70)
m3
Applying the thermal equation of state brings
m
ρ= (20.71)
V
so that the mass m finally follows

m = ρV = 1.205 kg (20.72)

• Approach 2:
The partial pressure of the air results from Dalton’s law

pa = p − pv = 996.7372 mbar. (20.73)

If the thermal equation of state is only applied to dry air, the following is
obtained
pa V
ma = = 1.1927 kg. (20.74)
Ra T

From approach 1 it is known that

xv = 0.010307 (20.75)

so it follows
m v = xv m a = 0.0123 kg. (20.76)

Finally, the entire mass is

m = m a + m v = 1.205 kg (20.77)

(b) Saturated state


Vapour is supplied to the closed system. Since the mass of dry air is constant,
the total pressure increases. For the partial pressure of the air this results in

m a Ra T
pa = = const. = 996.7372 mbar. (20.78)
V
The partial pressure of the vapour equals the saturated vapour pressure at 18 ◦ C,
i.e.
pv = ps (18 ◦ C) = 20.6466 mbar. (20.79)
20.2 Specific State Values 581

The maximum vapour load in saturated state is


pv pv
xv = 0.622 = 0.622 = 0.0128843. (20.80)
p − pv pa

Finally, the mass of the vapour in saturated state follows

m v = xv m a = 0.0154 kg (20.81)

The total pressure under these conditions is

p = pa + ps = 1017.4 mbar (20.82)

20.2.2 Caloric State Values

As this chapter is going to show, the specific enthalpy plays an important role in
technical applications. Therefore, the relevance of this specific state value is justified:
• In case, the system to be investigated is open, the first law of thermodynamics in
differential notation reads as

δq + δwt = dh + dea (20.83)

Obviously, the specific enthalpy is required to calculate open systems.


• Now the focus is on closed systems. The first law of thermodynamics in differential
notation obeys
δq + δw = du + dea . (20.84)

With the partial energy equation

δw = δwt + δwmech + δψ = − p dv + dea + δψ (20.85)

the first law of thermodynamics simplifies to

δq − p dv + δψ = du. (20.86)

Introducing the differential of the specific entropy

dh = du + p dv + v d p (20.87)

results in
δq + δψ = dh − v d p (20.88)

In the field of air conditioning technology, many changes of state occur at constant
pressure, i.e. d p = 0. Thus, it finally is
582 20 Humid Air

δq + δψ = dh. (20.89)

Thus, the specific enthalpy also plays an important role for closed systems. This is
the reason why the focus in this chapter is on specific enthalpy, e.g. with the h 1+x , x-
diagram according to Mollier, cf. Sect. 20.3. However, if the specific internal energy
u is required, e.g. for non-isobaric closed systems, it can easily be calculated by

u = h − pv. (20.90)

20.2.3 Specific Enthalpy h1+x

Since humid air is a multi-component mixture, i.e. it consists of water (vapour, liquid
and ice) and dry air, the total extensive enthalpy is composed as follows, see Sect.
19.3,

H = Ha + Hv + Hliq + Hice = m a · h a + m v · h v + m liq · h liq + m ice · h ice .


(20.91)
As mentioned above, the specific enthalpy h 1+x is related to the mass of dry air m a ,
i.e. it is
H mv m liq m ice
h 1+x = = ha + · hv + · h liq + · h ice . (20.92)
ma ma ma ma

Substituting the water content x leads to

h 1+x = h a + xv · h v + xliq · h liq + xice · h ice (20.93)

In this equation h a is the specific enthalpy of dry air and h w is the specific enthalpy
of water.21 Both are introduced in the following sections.

Specific enthalpy of dry air h a

Dry air is usually treated as an ideal gas. Thus, the specific enthalpy of dry air h a
within the multi-component system follows

dh a = c p,a · dT. (20.94)

The integration of the caloric equation of state yields22

21 So far, no distinction has been made in what aggregate state the water exists, i.e. it can be solid,
liquid or vapour.
22 The specific heat capacity of air c
p,a is considered constant.
20.2 Specific State Values 583

h a = c p,a · T − T0,a + h 0,a . (20.95)

Thus, a reference level h 0 T0,a is required and chosen23 as


h 0,a T0,a = 273.15 K = 0 (20.96)

This reference value is advantageous because it significantly simplifies the calculation


with the ◦ C scale, so that its handling is easy, i.e.

h a = c p,a · (T − 273.15 K) = c p,a · ϑ (20.97)

The specific isobaric heat capacity for dry air is

kJ
c p,a = 1.004 . (20.98)
kg K

Specific enthalpy of water h w

As known from Sect. 12.3, the caloric state value specific enthalpy of a real fluid is
a function of two independent state values, e.g. pressure p and temperature T , from
which follows24
dh w = dh w ( p, T ) . (20.99)

Therefore, analogous to the specific enthalpy of air, a reference value is required for
water. This reference level does not have to be the same as it has been for the compo-
nent air. Since for both components their specific mass balance must be fulfilled and
the first law of thermodynamics applies to the change of state, different reference
levels for water as for air do not lead to a dilemma.25 Following Eq. 20.99, in addition
to a reference temperature ϑ0,w , a reference pressure p0,w is required. Although the
choice is arbitrary, the definition for a reference level for water within the multicom-
ponent system follows the reference for the steam tables, see Sect. 18.3.7, i.e. the
saturated liquid state ( ) at the triple point is defined, thus

kJ
h 0,w (T0,w = 273.16 K, p0,w = 0.006117 bar) = 0.00061178 ≈0 (20.100)
kg

kJ
According to Table A.1, the specific enthalpy at reference level is 0.00061178 kg ,
since the specific internal energy at reference level has been set to zero, cf. Sect.
18.3.7, i.e.

23 This choice is arbitrary but practicable.


24 Assuming that the fluid is an ideal gas, i.e. the vapour, this equation simplifies to dh = dh (T ),
refer to Sect. 12.1.
25 Note that the reference values for water and for air must not be changed during a calculation.
584 20 Humid Air

Fig. 20.6 How to treat the water in humid air—vapour

u 0,w (ϑ0,w = 0.01 ◦ C, p0,w = 0.006117 bar) = 0. (20.101)

However, the specific reference enthalpy is pragmatically set to zero,26 cf. Eq. 20.100.
As already discussed, the water in the mixture of humid air can occur in different
aggregate states. To determine the thermodynamic state values of water, the steam
table of water could be referred to, but usually a more practical approach is preferred.
This approach is discussed by distinguishing between the following possible cases:
• Vapour
Commonly, the vapour occurring in humid air is considered an ideal gas. To deter-
mine its specific enthalpy, the approach is divided into two steps, cf. Fig. 20.6.
1. In the reference state, i.e. saturated liquid ( ) at the triple point (TP), the spe-
cific enthalpy of water is assumed to be zero, see Eq. 20.100. At first, energy
is supplied isobarically to vaporise the water.27 This is done until saturated
vapour at the triple point ( ) is reached. According to the steam table the heat
of vaporisation
kJ
h v (0.01 ◦ C) ≈ 2500 (20.102)
kg

is required. Thus, vapour is present after step 1.


2. From now on, the steam table approach is no longer applied, but the vapour is
treated as an ideal gas. Note that for ideal gases the specific enthalpy is solely
a function of temperature and thus there is no further pressure dependence. In
this second step, the vapour is heated or cooled to adjust the temperature of the
humid air28 from ϑ0,w to ϑ. In this step, the ideal gas vapour is subject to the
change in enthalpy of

h = c p,v T = c p,v · (ϑ − 0.01 ◦ C) ≈ c p,v · ϑ. (20.103)

Again, a pragmatic simplification is made by setting the reference temperature


at approximately 0 ◦ C for the temperature difference. The specific isobaric heat

26 Actually, this simplification is not a mistake, but a later comparison with the steam table, see
Tables A.1, A.2, A.3, A.4, A.5, A.6, A.7, A.8, A.9, A.10, A.11 and A.12, leads to minor deviations.
The choice of a reference level is always arbitrary.
27 The change of state from ( ) to ( ) shall be isobaric and thus also isothermal.
28 In most cases the humid air does not have a temperature of 0.01 ◦ C.
20.2 Specific State Values 585

capacity of vapour is
kJ
c p,v = 1.86 . (20.104)
kg K

Thus, the total specific enthalpy of vapour is the combination of steps 1 and 2:

kJ kJ
h v = h v (0.01 ◦ C) + c p,v · ϑ = 2500 + 1.86 ·ϑ (20.105)
      kg kg K
Step 1 Step 2

Example 20.3 The total pressure of humid air is p = 1 bar at a temperature of


ϑ = 20 ◦ C. The partial pressure of dry air is pa = 983 mbar. Thus, the partial pressure
of vapour is pv = 17 mbar. What is the specific enthalpy of the vapour?
– Approach 1 (Steam table—real fluid):

kJ
h v = h(20 ◦ C, 17 mbar) = 2537.8 (20.106)
kg

– Approach 2 (Ideal gas, according to Fig. 20.6):

kJ kJ kJ
h v = 2500 + 1.86 · ϑ = 2537.2 (20.107)
kg kg K kg

In other words, the error using the ideal gas approach compared to the real fluid
approach is significantly smaller than 0.1%. It turns out that using Eq. 20.105 is
easier to handle than applying the steam table.

• Liquid water
Starting from the reference point, see Fig. 20.7, which is already in the liquid state,
liquid water is now treated as an incompressible liquid.
According to Part I the change of enthalpy follows
 
h liq = h 0,w + cliq (ϑ − 0.01 ◦ C) +v p − p0,w ≈ cliq ϑ + v p − p0,w .
  
≈ϑ
(20.108)
Note that the pressure in the liquid p is equal to the pressure
 of the gaseous phase
in thermodynamic equilibrium. However, the term v p − p0,w is typically small

Fig. 20.7 How to treat the


water in humid air—liquid
586 20 Humid Air

and is usually ignored, so that the enthalpy of liquid water is as follows

h liq = cliq ϑ (20.109)

The specific heat capacity29 of liquid water is

kJ
cliq = 4.19 . (20.110)
kg K

Example 20.4 The total pressure of humid air is p = 1 bar at a temperature of


ϑ = 20 ◦ C. The humid air is already saturated and liquid water occurs. What is the
specific enthalpy of the liquid water30 ?
– Approach 1 (according to Eq. 20.10831 ):

h liq = cliq ϑ + v p − p0,w
kJ m3
= 4.19 · 20 K + 0.0010 (1 bar − 0.006117 bar) (20.111)
kg K kg
kJ
= 83.8994
kg

– Approach 2 (according to Eq. 20.109):

kJ
h liq = cliq ϑ = 83.8 (20.112)
kg

The influence of the term v p − p0,w is therefore rather small, the deviation
between the two approaches in this case is less than 0.12%.
– Approach 3 (according to Table A.7):

kJ
h liq = h(1 bar, 20 ◦ C) = 84.0118 (20.113)
kg

• Solid water
Starting from the reference point, see Fig. 20.8, which is in the liquid state, two
successive steps are required:
1. In the reference state, i.e. saturated liquid at the triple point (TP), the specific
enthalpy of water is assumed to be zero, see Eq. 20.100. At first, thermal energy

29 Note that for incompressible fluids there is no distinction between c p and cv .


30 Note that the liquid water has the same pressure as the gaseous phase in thermodynamic equilib-
rium.
31 In the equation, the temperature must be entered in ◦ C. However, since it is actually a temperature

difference compared to 0 ◦ C, 20 K is inserted in Eq. 20.111.


20.2 Specific State Values 587

Fig. 20.8 How to treat the water in humid air—ice

is released isobarically to freeze the water.32 This continues until the liquid has
completely turned to ice at the triple point. According to [1] the specific heat of
melting is
kJ
h m (0.01 ◦ C) ≈ −333 . (20.114)
kg

Thus, after step 1 ice at reference temperature is available.


2. From now on, the ice is treated as an incompressible solid. The change in
enthalpy for the ice in step 2 until the temperature ϑ is reached is therefore as
follows
 
h = cice (ϑ − 0.01 ◦ C) +v p − p0,w ≈ cice ϑ + v p − p0,w . (20.115)
  
≈ϑ

The specific heat capacity33 of ice is

kJ
cice = 2.05 . (20.116)
kg K

Note that the pressure within the solid p is equal to the pressure
 of the gaseous
phase in thermodynamic equilibrium. However, the term v p − p0,w is typi-
cally small and is usually ignored, so that the enthalpy of solid warer follows

h = cice ϑ. (20.117)

The total specific enthalpy of a solid, i.e. of ice, thus results from the combination
of steps 1 and 2, i.e.

kJ kJ
h ice = h m (0.01 ◦ C) + cice · ϑ = −333 + 2.05 ·ϑ (20.118)
      kg kg K
Step 1 Step 2

32 During this phase change, pressure and temperature remain constant.


33 Note that for incompressible solids there is no distinction between c p and cv .
588 20 Humid Air

20.2.4 Specific Entropy s1+x

Since humid air is a multi-component mixture, i.e. it consists of water (vapour, liquid
and ice) and dry air, the total extensive entropy is composed as follows, see Sect. 19.3:

S = Sa + Sv + Sliq + Sice = m a · sa + m v · sv + m liq · sliq + m ice · sice (20.119)

As discussed before, the specific entropy s1+x is referred to the mass of dry air m a ,
so that it is
S mv m liq m ice
s1+x = = sa + · sv + · sliq + · sice . (20.120)
ma ma ma ma

Substituting the water content x leads to

s1+x = sa + xv · sv + xliq · sliq + xice · sice (20.121)

In this equation sa is the specific enthalpy of dry air and sw the specific enthalpy of
water. Both are introduced in the following sections.

Specific entropy of dry air sa

Dry air can be treated as an ideal gas. However, similar to the specific enthalpy, see
Sect. 20.2.3, a reference level must be defined. Consequently, it is the same T0,a and
p0,a as before, i.e.

s0,a T0,a = 273.15 K, p0,a = 1 bar = 0 (20.122)

With this reference level the specific entropy for air is, see Part I,

T pa
sa − s0,a = c p,a ln − Ra ln. (20.123)
T0,a p0,a

Finally, it is
T pa
sa = c p,a ln − Ra ln (20.124)
T0,a p0,a

Specific entropy of water sw

Analogous to the specific entropy of air, a reference level is needed for water. As
already shown, this reference value does not have to be the same as for the component
air. Although the choice is arbitrary, the definition for a reference level for water
within the multi-component system follows the reference for the steam tables, see
Sect. 18.3.7, i.e. it is the saturated liquid state ( ) at the triple point, i.e.
20.2 Specific State Values 589

Fig. 20.9 How to treat the water in humid air—vapour

s0,w (T0,w = 273.16 K, p0,w = 0.006117 bar) = 0 (20.125)

As discussed earlier, the water in the mixture of humid air can occur in different states
of aggregation. To determine the thermodynamic state values of water, the steam table
of water could be consulted, but a more practical approach is usually preferred. This
approach is discussed by distinguishing between the following possible cases:
• Vapour
Commonly, the vapour occurring in humid air is considered to be an ideal gas. To
determine its specific entropy, the approach is divided into two steps
1. In the reference state, i.e. saturated liquid ( ) at the triple point (TP), the specific
entropy of water is set to zero, see Eq. 20.125. At first, thermal energy is supplied
isobarically to vaporise the water.34 This is done until saturated vapour ( ) at
the triple point is reached. According to the steam table the heat of vaporisation

kJ
h v (0.01 ◦ C) ≈ 2500 (20.126)
kg

is required. The supply of thermal energy is accompanied by the supply of


entropy, which is easy to calculate since the change of state is isothermal, i.e.

h v (0.01 ◦ C)
s = sa + si = . (20.127)
 T0,w
=0

2. In this second step, the ideal gas vapour is adjusted from ϑ0,w to ϑ respectively
from p0,w to pv .35 Thus, the caloric equation of state known from Part I can be
applied, i.e.
T pv
s = c p,v ln − Rv ln . (20.128)
T0,w p0,w

34 The change of state ( ) to ( ) shall be isobaric and thus isothermal as well. It is considered
reversible because there is no pressure loss, see Sect. 18.4.1, i.e. si = 0.
35 The vapour in the humid air possesses a temperature ϑ and its partial pressure is p .
v
590 20 Humid Air

The total specific entropy of the vapour therefore results from the combination of
steps 1 and 2, i.e.

h v (0.01 ◦ C) T pv
sv = + c p,v ln − Rv ln (20.129)
T0,w T0,w p0,w
     
Step 1 Step 2

Example 20.5 The total pressure of humid air is p = 1 bar at a temperature of


ϑ = 20 ◦ C. The partial pressure of dry air is pa = 983 mbar. Thus, the partial pressure
of vapour is pv = 17 mbar. What is the specific entropy of the vapour?
– Approach 1 (Steam table—real fluid):

kJ
sv = s (20 ◦ C, 17 mbar mbar) = 8.8145 (20.130)
kg K

– Approach 2 (Ideal gas, according to Fig. 20.9):

h v (0.01 ◦ C) T pv kJ
sv = + c p,v ln − Rv ln = 8.8118 (20.131)
T0,w T0,w p0,w kg K

This means that the error using the approach for ideal gas compared to the approach
for real fluid is significantly smaller than 0.1%.

• Liquid water
Starting from the reference point, see Fig. 20.10, which is already in the liquid
state, liquid water is now treated as an incompressible liquid. According to Part I
the change of entropy for an incompressible liquid follows

T T
sliq = s0,w + cliq ln = cliq ln (20.132)
T0,w T0,w

• Solid water
Starting from the reference point, see Fig. 20.11, which is in the liquid state, two
successive steps are required:

Fig. 20.10 How to treat the


water in humid air—liquid
20.2 Specific State Values 591

Fig. 20.11 How to treat the water in humid air—ice

1. In the reference state, i.e. saturated liquid at the triple point (TP), the specific
enthalpy of water is set to zero, see Eq. 20.100. At first, energy is released
isobarically to freeze the water.36 This continues until the liquid has completely
turned to ice at the triple point. The specific heat of melting is

kJ
h m (0.01 ◦ C) ≈ −333 . (20.133)
kg

The release of thermal energy is accompanied by the release of entropy,37 which


is easy to calculate because it is isothermal, i.e.

h m (0.01 ◦ C)
s = sa + si = < 0. (20.134)
 T0,w
=0

2. From now on, the ice is treated as an incompressible solid.38 The change in
entropy for the ice in step 2 until temperature ϑ is reached is therefore

T
sice = cice ln . (20.135)
T0,w

The total specific entropy of the solid, i.e. the ice, is consequently the combination
of steps 1 and 2, i.e.

h m (0.01 ◦ C) T
sice = + cice ln (20.136)
T0,w T0,w
     
Step 1 Step 2

36 During this phase change, pressure and temperature remain constant.


37 The change of state ( ) to ice shall be isobaric and thus isothermal as well. It is considered to run
reversible, since no pressure losses occur, see Sect. 18.4.1, i.e. si = 0.
38 This means that the specific entropy does not depend on pressure but only on temperature.
592 20 Humid Air

20.2.5 Overview Possible Cases

In this section an overview of possible states of humid air is given. Specific enthalpy
as well as specific entropy are listed based on the investigations of Sects. 20.2.3 and
20.2.4. If the supply of liquid or solid water is sufficiently large, evaporation takes
place until the gaseous phase is saturated with vapour, i.e. the system has reached a
thermodynamic equilibrium, see also Sect. 20.1.2. The required properties are listed
in Table 20.1, see [1].

• Case 1: Unsaturated air The water exists solely as vapour, see Fig. 20.12, so that
mw mv
x= = = xv . (20.137)
ma ma

Its specific enthalpy for this case is



h 1+x = c p,a · ϑ + x · h v + c p,v · ϑ (20.138)

The specific entropy obeys

Table 20.1 Properties—Humid air


Reference temperature air T0,a = 273.15 K
Reference temperature water T0,w = 273.16 K
Reference pressure air p0,a = 1 bar
Reference pressure water p0,w = 0.006117 bar
Gas constant dry air Ra = 287.0409 kgJK
Gas constant vapour Rv = 461.5158 J
kg K
Specific isobaric heat capacity air c p,a = 1.004 kgkJK
Specific isobaric heat capacity vapour c p,v = 1.86 kgkJK
Specific heat capacity liquid water cliq = 4.19 kgkJK
Specific heat capacity solid water cice = 2.05 kgkJK
Enthalpy of vaporisation at T0,w h v = h v (0.01 ◦ C) = 2500 kg
kJ

Enthalpy of melting at T0,w h m = h m (0.01 ◦ C) = −333 kg kJ

Fig. 20.12 Unsaturated


humid air
20.2 Specific State Values 593

Fig. 20.13 Saturated humid


air

 
T pa T pv h v
s1+x = c p,a ln − Ra ln + x c p,v ln − Rv ln +
T0,a p0,a T0,w p0,w T0,w
(20.139)
• Case 2: Saturated air
According to Fig. 20.13 the vapour load reaches its maximum, i.e.
mw mv
x= = = xv = x  . (20.140)
ma ma

The specific enthalpy for this case is



h 1+x = c p,a · ϑ + x  · h v + c p,v · ϑ (20.141)

The specific entropy obeys

 T p − pv
s1+x = c p,a ln − Ra ln + (20.142)
T0,a p0,a
 
T p h v,0
+ x  c p,v ln − Rv ln v + . (20.143)
T0,w p0,w T0,a

• Case 3: Saturated air + liquid water, i.e. ϑ > 0


In this case, sufficient water is available and is present as vapour and as liquid
water,39 see Fig. 20.14, so that

mw m v + m liq
x= = = xv + xliq = x  + xliq . (20.144)
ma ma

Its specific enthalpy for this case is

39 Thus, according to Fig. 18.9a, the temperature is ϑ > 0. In fact, the temperature dependence of

water freezing is rather small. Liquid water has the same pressure as the gaseous phase, so that
under atmospheric conditions it is almost 0 ◦ C at the beginning of freezing.
594 20 Humid Air

Fig. 20.14 Saturated humid


air and liquid water

Fig. 20.15 Saturated humid


air and solid water

 
h 1+x = h 1+x + xliq h liq = c p,a · ϑ + x  · h v + c p,v · ϑ + x − x  cliq ϑ
(20.145)

The specific entropy40 is

 
 T
s1+x = s1+x + xliq sliq = s1+x + x − x  cliq ln (20.146)
T0,w

The volume of liquid water is usually neglected compared to the volume of the
gaseous phase.
• Case 4: Saturated air + solid water, i.e. ϑ < 0
In this case, sufficient water is available and is present as vapour and as solid water,
see Fig. 20.15, so that

mw m v + m ice
x= = = xv + xice = x  + xice . (20.147)
ma ma

Its specific enthalpy for this case is

h 1+x = h 1+x + xice h ice (20.148)

respectively

40 Note that in thermodynamic equilibrium the pressure within the liquid is equal to the pressure of
the gaseous phase.
20.2 Specific State Values 595

Fig. 20.16 Saturated humid


air, liquid and solid water

 
h 1+x = c p,a · ϑ + x  · h v + c p,v · ϑ + x − x  (h m + cice · ϑ) (20.149)

The specific entropy41 is




 
 h m
 T
s1+x = s1+x + xice sice = s1+x + x−x + cice ln (20.150)
T0,w T0,w

The volume of solid water is usually neglected compared to the volume of the
gaseous phase.
• Case 5: Saturated air + liquid water + solid water, i.e. ϑ = 0
In this case, sufficient water is present and exists as vapour, as liquid water and as
solid water, see Fig. 20.16. This is under atmospheric conditions possible, when
the temperature is ϑ = 0 ◦ C, i.e. a three-phase mixture exists,42 so that

mw m v + m liq + m ice
x= = = xv + xliq + xice = x  + xliq + xice . (20.151)
ma ma

Its specific enthalpy for this case is

h 1+x = h 1+x + xliq h liq + xice h ice (20.152)

The specific entropy43 is


s1+x = s1+x + xliq sliq + xice sice (20.153)

The volume of the liquid as well as the solid water is usually neglected compared
to the volume of the gaseous phase.

41 Note that in thermodynamic equilibrium the pressure within the solid is equal to the pressure of
the gaseous phase.
42 See also Fig. 20.20. This state is not defined unambiguously and additional information is

required, e.g. about the ratio of liquid and ice.


43 Note that the pressure within the solid and liquid is equal to the pressure of the gaseous phase.
596 20 Humid Air

Fig. 20.17 Sketch to Problem 20.2

Problem 20.2 A cylinder contains V1 = 5 m3 air as well as 0.119 kg liquid water


in thermodynamic equilibrium (1), cf. Fig. 20.17. The volume of the liquid water
shall be ignored. The total pressure is p1 = 2.0 bar at a temperature of ϑ1 = 25 ◦ C.
The weights on top of the piston, that closes the cylinder tightly, are removed very
slowly and step by step, so that the humid air expands isothermally and without any
dissipation until state (2) is reached. In this state the liquid water is fully vaporised.
Air and vapour can be treated as ideal gases (Ra = 0.287 kgkJK , Rv = 0.4615 kgkJK ). The
potential energy of the cylinder content can be disregarded. Ambient state is ϑenv =
ϑ1 and penv = 0.8 bar. System and environment are perfectly thermally coupled.
(a) What is the mass of the dry air and the mass of the entire water within the
cylinder?
(b) Calculate volume V2 after the expansion? What is the pressure p2 ?
(c) Calculate the released work W12 during the expansion.
(d) What amount of heat Q 12 is exchanged with the environment?
(e) What is the change of entropy of the cylinder content during the process?
(f) What is the change of exergy E x = E x,2 − E x,1 of the closed system?
Solution
The following Fig. 20.18 shows the change of state and the notation of the relevant
variables to solve this task. Since the system in state (1) is in thermodynamic equi-
librium and the amount of liquid water is sufficiently large, the air is saturated with
vapour.
(a) Because the air is saturated, the partial pressure of the vapour can be taken from
the steam table, i.e.
pv1 = ps (ϑ1 ) = 0.0317 bar. (20.154)

According to Dalton, the partial pressure of the air is then obtained as follows
20.2 Specific State Values 597

Fig. 20.18 Solution to Problem 20.2

pa1 = p1 − pa1 = 1.9683 bar. (20.155)

The application of the thermal equation for the air results in

pa1 V1
m a1 = = 11.5013 kg = const. = m a2 = m a . (20.156)
Ra T1

The mass of the vapour follows accordingly

pv1 V1
m v1 = = 0.1152 kg. (20.157)
Rv T1

Hence, the entire mass of the water is

m w1 = m v1 + m liq1 = 0.2342 kg = const. = m w2 = m w (20.158)

(b) In state (2), the liquid water has just completely turned into vapour, so that state
(2) is saturated state. Thus, the partial pressure of the vapour44 is

pv2 = ps (ϑ2 = ϑ1 ) = 0.0317 bar = pv1 = pv . (20.159)

The entire water now is turned into vapour, i.e.

m v2 = m w1 = 0.2342 kg. (20.160)

44 See steam table.


598 20 Humid Air

Application of the thermal equation of state for the vapour in state (2), i.e.

m v2 Rv T2
V2 = = 10.1657 m3 (20.161)
pv2

The mass of the dry air remains constant, so that

m a1 Ra T2
pa2 = = 0.9681 bar. (20.162)
V2

The total pressure then follows

p2 = pa2 + pv2 = 0.9998 bar (20.163)

(c) The partial energy equation for the expansion reads as

W12 = W12,V + W12,mech + 12 . (20.164)


   
=0 =0

It is known that the volume work W12,V causes for the volume change. In this
case, the total volume of the cylinder at any point in time is composed of the
volume of the gaseous phase and the volume of the liquid phase. During the
change of state, the volume of the liquid phase shrinks to zero, while the volume
of the gaseous phase increases. Accordingly, the volume work can be split into
two parts,45 see Fig. 20.18:
• gaseous phase V1,gas → V2 with p1 → p2
– air V1,gas → V2 with pa1 → pa2
– vapour V1,gas → V2 with pv1 = pv2
• liquid phase Vliq1 → 0 with pliq1 = p1 → pv2 . Note that Vliq1 ≈ 0.
Hence, the work W12 follows

W12 = W12,V = W12,V,gas + W12,V,liq . (20.165)

Introducing the volume work and applying Dalton’s law results in

45 It is the changing volume, that causes work, since δWV = − p A dx = − p dV . Thus, the mass
is not involved directly. However, the mass is included due to the thermal equation of state. Con-
sequently, it is the volume change of the gaseous phase, i.e. air and vapour, as well as the volume
change of the liquid, that needs to be taken into account. The mass of the vaporising liquid is the
reason why the partial pressure of the vapour in the gaseous phase is constant.
20.2 Specific State Values 599

V2 0 V2
W12 = − p dVgas − pliq dVliq = − ( pa + pv ) dVgas − 0
V1,gas Vliq1 V1,gas
     
gaseous phase =0, since Vliq1 ≈0 (20.166)
V2 V2
=− pa dVgas − pv dVgas .
V1,gas V1,gas

As mentioned above, the volume of the liquid can be ignored, so that V1,gas = V1 .
Since the partial pressure of the vapour is constant, one gets

V2
W12 = − pa dVgas − pv (V2 − V1 ) . (20.167)
V1

Applying the isothermal volume work for air, treated as ideal gas, results in

pa1
W12 = −m a Ra T1 ln − pv (V2 − V1 ) = −7.1471 × 105 J (20.168)
pa2

(d) The first law of thermodynamics leads to the exchanged heat Q 12 . The change
of the internal energy is due to the change of the internal energy of the air m a ,
the change of the internal energy of the initial vapour m v1 and the change of the
internal energy of the liquid m liq1 , see also Fig. 20.19, i.e.

W12 + Q 12 = Utotal = Ua + Uv + Uliq . (20.169)

The application of the caloric equation of state, treating air and vapour as ideal
gas, leads to

Fig. 20.19 Solution to


Problem 20.2—Phase
change of the water
600 20 Humid Air

W12 + Q 12 = Utotal
(20.170)
= m a cv,a (T2 − T1 ) + m v1 cv,v (T2 − T1 ) + m liq1 u.

Since, the change of state is isothermal one gets



Q 12 = m liq1 u − W12 = m liq1 u v2 − u liq1 −W12 . (20.171)
  
=u

With
kJ V2 kJ
u v2 = h v2 − pv2 vv2 = 2500 + c p,v ϑ2 − pv2 = 2408.9 (20.172)
kg mw kg

respectively

kJ
u liq1 = h liq1 − p1 vliq1 = cliq ϑ1 − p1 v ( p1 , T1 ) = 104.5494 (20.173)
   kg
Steam table

one finally gets



Q 12 = m liq1 u − W12 = m liq1 u v2 − u liq1 − W12 = 9.8893 × 105 J
(20.174)

(e) There are two alternatives in order to calculate the change of entropy:
• Alternative 1: Process values
The change of entropy for this reversible process follows

S = S2 − S1 = Si,12 +Sa,12 . (20.175)



=0

Since the change of state is isothermal, it simplifies to

Q 12 J
S = S2 − S1 = Sa,12 = = 3.317 × 103 (20.176)
T1 K

• Alternative 2: State values


The total entropy change is due to the entropy change of the air, vapour and
liquid, corresponding to the internal energy in part (d), i.e.

S = Sa + Sv + Sliq (20.177)

with
20.2 Specific State Values 601

– Air

Sa = Sa2 − Sa1





T2 pa2 T1 pa1
= m a c p,a ln − Ra ln − c p,a ln − Ra ln
T0,a p0,a T0,a p0,a
J
= 2.3422 × 103
K
(20.178)
– Vapour

Sv = Sv2 − Sv1


 
h v T2 pv2
= m v1 + c p,v ln − Rv ln +
T0,w T0,w p0,w
  (20.179)
h v T1 pv1
− m v1 + c p,v ln − Rv ln
T0,w T0,w p0,w
=0

This is since the temperature and the partial pressure of the vapour remain
constant.
– Liquid

Sliq = Sliq2 − Sliq1




h v T2 pv T1
= m liq1 + c p,v ln − Rv ln 2 − cliq ln
T0,w T0,w p0,w T0,w
J
= 974.4833 .
K
(20.180)
Thus, the entire change of entropy results in

J
S = Sa + Sv + Sliq = 3.317 × 1063 (20.181)
K

The deviation of approach 1 and 2 is less than 0.01% and is most likely due to
inaccuracies in the properties of the fluids with respect to the steam table.
(f) The balance of exergy leads to

E x,2 = E x,1 + Weff + E x,Q − E x,V . (20.182)

The heat transfer takes place at ambient temperature, so that

E x,Q = 0. (20.183)
602 20 Humid Air

Furthermore, the change of state is reversible, i.e.

E x,V = 0. (20.184)

One therefore obtains

E x = E x,2 − E x,1 = Weff = W12 + penv (V2 − V1 ) = −3.0145 × 105 J


(20.185)

20.3 The h1+x , x-diagram According to Mollier

Obviously, humid air is a two- respectively three-phase mixture and the relevant
correlations to describe its thermodynamic states have been introduced in the previous
sections. It has been clarified under which conditions liquid or even solid water occurs
in humid air. In order to deal with changes of state, it is advantageous to visualise
states of equilibrium, e.g. in p, v- or T, s-diagrams. This has been done in Part I,
so that process values for instance can be easily illustrated. Especially in the field
of HVAC technology, the application of the so-called h 1+x , x-diagram according to
Mollier,46 which is presented in this section, is advantageous. A schematic h 1+x , x-
diagram is shown in Fig. 20.20. A detailed diagram can be found in Appendix D:
It includes a h 1+x , x-diagram for a mixture of atmospheric47 air and water, cf. Fig.
D.1. Figure D.2 further shows a h 1+x , x-diagram for a hydrogen/water mixture, that
may be useful for fuel cell applications in order to handle the humidification of the
hydrogen. Anyhow, all correlations discussed so far are gathered and visualised in
such diagrams. The water content x is plotted on the x-axis, the specific enthalpy
h 1+x is shown on the y-axis. Figure 20.20 visualises the regions previously discussed,
i.e. it shows the state of the water in the humid air. Characteristic for the h 1+x , x-
diagram is the so-called saturation curve, i.e. the curve with a relative humidity of
ϕ = 1. Above that curve, the humid air is unsaturated, so that the relative humidity ϕ
is less than 100%. Below the ϕ = 1 curve, the humid air is saturated and additional
liquid or solid water appears. The following examples summarise possible states in
thermodynamic equilibrium:
• State (1)
Unsaturated humid air, single-phase region, i.e. vapour and dry air are in gaseous
state
• State (2)
Saturated humid air, the humid air has its maximum vapour load with respect to
its temperature

46 Richard Mollier (30 November 1863 in Triest, 13 March 1935 in Dresden). At the 1923
Thermodynamics Conference held in Los Angeles it was decided to name all state diagrams, that
have the specific enthalpy h as one of its axis, after Mollier.
47 Composition: ξ
N2 = 0.77, ξO2 = 0.23.
20.3 The h 1+x , x-diagram According to Mollier 603

Fig. 20.20 Mollier h 1+x , x-diagram—schematic

• State (3)
Saturated humid air (gaseous phase composed of dry air and vapour) and ice (solid
phase), i.e. two-phase region
• State (4)
Saturated humid air (gaseous phase composed of dry air and vapour) and ice (solid
phase) and water (liquid phase), i.e. three-phase region
• State (5)
Saturated humid air (gaseous phase composed of dry air and vapour) and water
(liquid phase), i.e. two-phase region
However, there are some features of the h 1+x , x-diagram that require further expla-
nation:
• Lines of constant enthalpy h 1+x are inclined so that the ϑ = 0 ◦ C isotherm for
unsaturated air, i.e. in the single-phase region, runs horizontally, see Fig. 20.20.
However, the other isotherms in the unsaturated range are not perfectly horizontal.
• At the saturation line (ϕ = 1) the isotherms kink, so that their gradient becomes
significant smaller than in the unsaturated region:

– Single-phase region
  
∂h 1+x,1P ∂ c p,a · ϑ + x · h v + c p,v · ϑ

=
∂ x ϑ=const. ∂x
ϑ (20.186)
= h v + c p,v · ϑ
604 20 Humid Air

– Two-phase region (liquid)


   
∂h 1+x,2P ∂ c p,a ϑ + x  h v + c p,v ϑ + x − x  cliq ϑ

=
∂ x ϑ=const. ∂x
ϑ
= cliq · ϑ
(20.187)

It obviously follows that



∂h 1+x,2P ∂h 1+x,1P
< (20.188)
∂x ϑ=const. ∂x ϑ=const.

Therefore, the isotherms kink at the saturation curve. However, lines of constant
enthalpy h 1+x run through the different regions without kinks.
• As mentioned before, three phases can occur at ϑ = 0 ◦ C, i.e. gaseous, liquid
and solid phase. Hence, this isotherm splits at the saturation curve ϕ = 1 into
liquidus and solidus line, see Fig. 20.20. According to Eq. 20.187, the gradient of
the liquidus line is
∂h 1+x
= cliq · ϑ = 0 (20.189)
∂ x ϑ=0 ◦ C

The liquidus line thus runs congruent to its isenthalp.

For the solidus line it is


   
∂h 1+x ∂ c p,a ϑ + x  h v + c p,v ϑ + x − x  (h m + cice ϑ)

=
∂ x ϑ=0 ◦ C ∂x ◦
ϑ=0 C
= h m + cice · ϑ = h m
<0
(20.190)
• Usually the h 1+x , x-diagram is given for a total pressure p = 1 bar. A conversion
to other pressures is given later, see Sect. 20.4.6.

20.4 Changes of State for Humid Air

The handling of the h 1+x , x-diagram is demonstrated in this section for changes of
state of humid air relevant for ventilation and air conditioning. The focus is on the
thermodynamic balances on the one hand. These balances include the conservation
of mass for water and air as well as the first law of thermodynamics. Each change of
state is further illustrated on the other side in the h 1+x , x-diagram so that its benefit
becomes evident.
20.4 Changes of State for Humid Air 605

Fig. 20.21 Heating (1) → (2) and cooling (1) → (3) at constant water content

20.4.1 Heating and Cooling at Constant Water Content

This change of state is exemplified in Fig. 20.21. As the entire water content remains
constant,48 i.e. x = const., the change of state follows a vertical line. In order to
perform the relevant balances, a distinction is made between open and closed systems:
• Open system
The first law of thermodynamics49 in steady state obeys

Q̇ 12 = Ḣ2 − Ḣ1 . (20.191)

Introducing the specific enthalpy of humid air h 1+x,2 , that is referred to the mass
of the dry air, brings

Q̇ 12 = ṁ a · h 1+x,2 − h 1+x,1 (20.192)

Thus, a heating, i.e. Q̇ > 0 leads to an increase of specific enthalpy, a cooling, i.e.
Q̇ < 0 to a decrease of the specific enthalpy.

48 This is due to the fact that no water is added or removed. Only thermal energy is supplied or
released, respectively.
49 Kinetic as well as potential energy are ignored.
606 20 Humid Air

The mass conservation for the water is

ṁ 1,w = ṁ 2,w = ṁ w (20.193)

respectively for the air


ṁ 1,a = ṁ 2,a = ṁ a . (20.194)

With the entire mass


ṁ = ṁ a + ṁ w = const. (20.195)

so that

ṁ a = (20.196)
1+x

• Closed system
The first law of thermodynamics in differential notation50 is

δ Q + δW = dU. (20.197)

Applying the partial energy equation51

δW = − p dV (20.198)

and the definition for the enthalpy brings

δ Q − p dV = dH − p dV − V d p. (20.199)

As already mentioned, in HVAC applications mostly isobaric changes of state


occur, i.e. d p = 0, so that
Q 12 = m · (h 2 − h 1 ) (20.200)

Introducing the specific enthalpy for humid air h 1+x,2 , i.e.



Q 12 = m a · h 1+x,2 − h 1+x,1 (20.201)

The mass conservation for the water is

m 1,w = m 2,w = m w (20.202)

50 Let us assume, the change of state is reversible and kinetic as well as potential energy can be
ignored.
51 No mechanical work, since kinetic and potential energy are ignored.
20.4 Changes of State for Humid Air 607

respectively for the air


m 1,a = m 2,a = m a . (20.203)

With the entire mass


m = m a + m w = const. (20.204)

so that
m
ma = (20.205)
1+x

Obviously, according to Fig. 20.21 two options are possible:


• Heating (1) → (2)
– Water content x remains constant
– Enthalpy h 1+x increases, see first law of thermodynamics
– Temperature ϑ increases
– Relative humidity ϕ decreases
• Cooling (1) → (3)
– Water content x remains constant
– Enthalpy h 1+x decreases, see first law of thermodynamics
– Temperature ϑ decreases
– Relative humidity ϕ increases
– State (3) may even be located in the two-phase region in case cooling is sufficient

20.4.2 Dehumidification

This change of state is essential in air conditioning systems, e.g. for operation in
summer. As already known, air at high temperature potentially contains a large
amount of water vapour. Therefore, it may be necessary to dehumidify the air to
prevent the condensation of water when the air is cooled. This example is shown in
Fig. 20.22 for an open system at steady state. It is obvious that three components are
required to design an air conditioning system: a chiller,52 a water separator and a
heater:
• In the chiller, the humid air (1) is cooled, while the water content x remains
constant, as no water is released or supplied. The change of state thus runs vertically
downwards in the h 1+x , x diagram. In state (1 ), the humid air reaches the saturated
state. A further release of thermal energy leads to liquid water, while the gaseous
phase remains saturated, i.e. the two-phase range is reached and cooling ends at
state (2).

52 A chiller is a cooler.
608 20 Humid Air

Fig. 20.22 Dehumidification of air

The first law of thermodynamics53 obeys



Q̇ 12 = Ḣ2 − Ḣ1 = ṁ a · h 1+x,2 − h 1+x,1 (20.206)

with the specific enthalpies



h 1+x,1 = c p,a ϑ1 + x1 · h v + c p,v ϑ1 (20.207)

respectively, since the air with a temperature of ϑ2 can only carry x2 as vapour
load

h 1+x,2 = c p,a ϑ2 + x2 · h v + c p,v ϑ2 + (x2 − x2 ) · cliq · ϑ2 . (20.208)

The mass balance for dry reads as

ṁ a = const. = ṁ 1,a = ṁ 2,a . (20.209)

The mass balance for water reads as

53 Let us assume, kinetic as well as potential energy can be ignored.


20.4 Changes of State for Humid Air 609

ṁ w = const. = ṁ 1,w = ṁ 2,w . (20.210)

In state (1) the water is in vapour state, i.e.

ṁ 1,w = ṁ 1,v = x1 ṁ a . (20.211)

In state (2) it occurs as vapour54 and liquid

ṁ 2,w = ṁ 2,v + ṁ 2,liq = x2 ṁ a + ṁ 2,liq . (20.212)

Thus, the mass of the liquid water results in



ṁ 2,liq = x1 ṁ a − x2 ṁ a = ṁ a x1 − x2 = ṁ a x. (20.213)

• In the second step, the liquid water is released by a separator. Let us assume, the
separator works adiabatically, so the first law of thermodynamics yields

ṁ a h 1+x,2 = ṁ a h 1+x,2 + ṁ 2,liq h 2,liq . (20.214)

Division by ṁ a results in

h 1+x,2 = h 1+x,2 + x h 2,liq (20.215)

with
h 2,liq = cliq · ϑ2 (20.216)

and 
h 1+x,2 = c p,a ϑ2 + x2 · h v + c p,v ϑ2 . (20.217)

A combination of Eqs. 20.215, 20.216, 20.217 and 20.208 brings

ϑ2 = ϑ2 . (20.218)

The water separator therefore works isothermally.


• Now the dehumidified air is heated up again until ϑ3 = ϑ1 is reached. The required
thermal energy can be calculated according to the first law of thermodynamics,
i.e. 
Q̇ 2 3 = ṁ · (h 3 − h 2 ) = ṁ a · h 1+x,3 − h 1+x,2 (20.219)

with the specific enthalpy of state (3)



h 1+x,3 = c p,a ϑ1 + x2 · h v + c p,v ϑ1 . (20.220)

54 In saturated state.
610 20 Humid Air

Fig. 20.23 Adiabatic mixing of humid air

According to Fig. 20.22, state (3) finally has the same temperature as the initial state
(1), but the air has been dehumidified. The relative humidity ϕ and the water content
x have been reduced. If further dehumidification is desired, the air needs a further
temperature reduction inside the cooler.

20.4.3 Adiabatic Mixing of Humid Air

In this section, the adiabatic mixing of humid air is investigated. This change of state
is important when an air conditioner is operated in bypass mode, for example, i.e.
fresh air is mixed with recirculated air. The principle is illustrated in Fig. 20.23: One
air flow (1) is mixed adiabatically with a second flow (2) and the mixed air (3) leaves
the system boundary. As already mentioned, a mass balance is required for both
the dry air and the water, which follows the idea of mass conservation. In addition,
however, an energy balance is also required. First, the mathematical relationship is
derived. In the second step, the change of state is represented in an h 1+x , x-diagram.
Let us start with the mass balance for dry air, i.e.

ṁ a,3 = ṁ a,1 + ṁ a,2 . (20.221)

The mass conservation for water55 obeys

ṁ w,3 = ṁ w,1 + ṁ w,2 . (20.222)

With introducing the water content x one gets:

ṁ a,3 x3 = ṁ a,1 x1 + ṁ a,2 x2 . (20.223)

These equations can be solved for the water content of the humid mixed air x3 , i.e.

55 At this stage no distinction is made regarding the state of the water.


20.4 Changes of State for Humid Air 611

Fig. 20.24 Adiabatic mixing of humid air, h 1+x , x-diagram

ṁ a,1 x1 + ṁ a,2 x2 ṁ a,1 x1 + ṁ a,2 x2


x3 = = (20.224)
ṁ a,3 ṁ a,1 + ṁ a,2

Once the water content x3 is known, the following equations can be easily derived,
i.e. 
x1 ṁ a,1 + ṁ a,2 − ṁ a,1 x1 − ṁ a,2 x2
x1 − x3 =
ṁ a,1 + ṁ a,2
(20.225)
ṁ a,2 (x1 − x2 )
=
ṁ a,1 + ṁ a,2

respectively 
ṁ a,1 x1 + ṁ a,2 x2 − x2 ṁ a,1 + ṁ a,2
x3 − x2 =
ṁ a,1 + ṁ a,2
(20.226)
ṁ a,1 (x1 − x2 )
= .
ṁ a,1 + ṁ a,2

Thus it is finally
x1 − x3 ṁ a,2 a
= = (20.227)
x3 − x2 ṁ a,1 b

The distances a and b are depicted in the h 1+x , x diagram, see Fig. 20.24. The first
law of thermodynamics for the adiabatic mixing in steady state56 is applied, i.e.

56This means that energy flux in is balanced by the energy flux out. Kinetic as well as potential
energies are neglected.
612 20 Humid Air

Ḣ1 + Ḣ2 = Ḣ3 . (20.228)

Introducing the specific enthalpy leads to



ṁ a,1 h 1+x,1 + ṁ a,2 h 1+x,2 = ṁ a,1 + ṁ a,2 h 1+x,3 . (20.229)

Hence, the specific enthalpy of the mixed air (3) follows

ṁ a,1 h 1+x,1 + ṁ a,2 h 1+x,2


h 1+x,3 = (20.230)
ṁ a,1 + ṁ a,2

Once the specific enthalpy h 1+x,3 is known, the following equations can be easily
derived

h 1+x,1 ṁ a,1 + ṁ a,2 − ṁ a,1 h 1+x,1 − ṁ a,2 h 1+x,2
h 1+x,1 − h 1+x,3 =
ṁ a,1 + ṁ a,2
 (20.231)
ṁ a,2 h 1+x,1 − h 1+x,2
=
ṁ a,1 + ṁ a,2

respectively

ṁ a,1 h 1+x,1 + ṁ a,2 h 1+x,2 − h 1+x,2 ṁ a,1 + ṁ a,2
h 1+x,3 − h 1+x,2 =
ṁ a,1 + ṁ a,2
 (20.232)
ṁ a,1 h 1+x,1 − h 1+x,2
= .
ṁ a,1 + ṁ a,2

Hence, it finally is
h 1+x,1 − h 1+x,3 ṁ a,2 c
= = (20.233)
h 1+x,3 − h 1+x,2 ṁ a,1 d

The distances c and d are depicted in the h 1+x , x-diagram,57 cf. Fig. 20.24.

A comparison of Eqs. 20.227 and 20.233 leads to

h 1+x,1 − h 1+x,3 x1 − x3
= . (20.234)
h 1+x,3 − h 1+x,2 x3 − x2

Rearranging this equation brings

h 1+x,3 − h 1+x,1 h 1+x,2 − h 1+x,3


= . (20.235)
x3 − x1 x2 − x3

57Note that it has not yet been proved that (1), (2) and (3) lie on a straight line. The proof follows
now.
20.4 Changes of State for Humid Air 613

Fig. 20.25 Adiabatic mixing of humid air, h 1+x , x-diagram

These terms represent the gradients of the curves 31 and 23:



h c d h
= = = (20.236)
x 31 a b x 23

The gradients in the inclined h 1+x , x-diagram are therefore identical. Consequently,
points (1), (2) and (3) must lie on a straight line. Figure 20.25 summarises the change
of state for an adiabatic mixing of two flows. Equation 20.227 can be extended: Since
all points are located on one straight line, it also follows,58 see Fig. 20.25:

ṁ a,2 x1 − x3 13
= = (20.237)
ṁ a,1 x3 − x2 23

Theorem 20.1 A mass flux (1) of humid air is adiabatically mixed with a mass flux
(2) of humid air. The new state (3) lies on a straight line between (1) and (2). This line
is called mixing straight. The greater the amount of dry air, the smaller the distance
to the mixing point:
• ṁ a,1 = ṁ a,2 : The mixing state (3) is located right in the middle of the line 12.

58 This is due to (x3 − x2 ) = 23 · cos α and (x1 − x3 ) = 13 · cos α.


614 20 Humid Air

Fig. 20.26 Humidification of air with pure water

• ṁ a,1 > ṁ a,2 : The mixing state (3) moves closer to (1).
• ṁ a,1 < ṁ a,2 : The mixing state (3) moves closer to (2).
Note that no restrictions have been placed on the region of the mixing states.
Thus, this principle also works if the mixing states are in the two/three phase region.
It is even possible that the mixing of two unsaturated streams of humid air (1 ) and
(2 ) leads to a state where liquid water (3 ) occurs, as shown in Fig. 20.25. This can
be observed in winter when the ambient air is cold and contains small amounts of
water vapour, see state (2 ). When a person exhales, this state tends to be warm and
vaporous, see state (1 ). The mixing then often leads to condensation of the breathing
air, see state (3 ).

20.4.4 Humidification of Air

Section 20.4.2 has shown that dehumidification of the air may be necessary, e.g. in
summer when the air potentially contains a large amount of water vapour due to its
temperature. However, in winter, which is characterised by cold and dry air, one of
the main tasks of HVAC systems is to humidify the air. This can be done with liquid,
solid or vapour water, which is injected into the rather dry air. The consequences are
discussed thermodynamically in this section. However, the principle is illustrated in
Fig. 20.26. The air that needs to be humidified shall be in state (1). The flow of pure
water59 is state (2) and the humidified air is state (3). Unfortunately, due to

ṁ w
x2 = →∞ (20.238)
ṁ a,2

the mixing-straight-approach, see Sect. 20.4.3, does not work any more.60 However,
the mass balance for air is

59 It is assumed that the water is supplied pure, i.e. without entrainment of dry air, i.e. ṁ a,2 = 0.
60 Due to x2 → ∞ state (2) can not be fixed in the h 1+x , x-diagram.
20.4 Changes of State for Humid Air 615

ṁ a,3 = ṁ a,1 = ṁ a = const. (20.239)

The mass balance for water reads as

ṁ w,3 = ṁ w,1 + ṁ w . (20.240)

Introducing the water content x brings

ṁ a x3 = ṁ a x1 + ṁ w . (20.241)

Hence, rearranging leads to

ṁ w
x3 − x1 = x = (20.242)
ṁ a

Application of the first law of thermodynamics in steady state61 brings

ṁ a h 1+x,1 + ṁ w h w = ṁ a h 1+x,3 . (20.243)

This equation can be rearranged, so that

ṁ w
h 1+x,3 − h 1+x,1 = h 1+x = hw (20.244)
ṁ a

Division of Eqs. 20.242 and 20.244 leads to

h 1+x h 1+x,3 − h 1+x,1


= = hw (20.245)
x x3 − x1

Obviously, this gradient hx1+x shows the direction a change of state takes in the
h 1+x , x-diagram. According to Eq. 20.245 this gradient is identical with the specific
enthalpy of the water that is injected. Starting from the initial state (1) the increase of
the water content x can be calculated according to Eq. 20.242. Therefore, the mass
of the supplied water ṁ w must be known. However, the knowledge of x is not
sufficient to determine state (3), since it can be at any vertical position corresponding
to the new water content
x3 = x1 + x. (20.246)

The second information to fix state (3) comes from the gradient information.
Figure 20.27 shows how this information can be applied in a h 1+x , x-diagram. Start-

61Both kinetic and potential energy are ignored. The incoming energy is balanced by the outgoing
energy.
616 20 Humid Air

Fig. 20.27 Humidification of air with pure water

ing from a so-called pole, the hx1+x -information can be tapped with a ruler along the
edges of the h 1+x , x-diagram. After the gradient has been defined by scale and pole,
it can be shifted in parallel by state (1). The point of intersection with the vertical
line containing the x-information results in the new state (3).

Example 20.6 The following examples show how to handle the h 1+x , x-diagram
for air humidification. Let us assume the initial state (1) is well known and can be
fixed in the h 1+x , x-diagram. Each of the following examples can be solved mathe-
matically by applying the derived correlations. However, the focus is on a graphical
solution. In most cases, this method is advantageous because the change of state can
be investigated quickly compared to a mathematical solution. The following steps
are required:
• Step 1: Determine x
This can be done easily by applying

ṁ w
x = → x3 = x1 + x. (20.247)
ṁ a

The vertical x3 -line can therefore be drawn in the h 1+x , x-diagram.


• Step 2: Determine the gradient hx1+x
Depending on the state of water the gradient follows:
20.4 Changes of State for Humid Air 617

– Case (a): Vapour (1) → (3)

h 1+x
= h w = h v = h v + c p,v · ϑ (20.248)
x

– Case (b): Liquid water (1) → (3 )

h 1+x
= h w = h liq = cliq · ϑ (20.249)
x

– Case (c): Soild water, i.e. ice (1) → (3 )

h 1+x
= h w = h ice = h m + cice · ϑ < 0 (20.250)
x
• Step 3: Graphical construction in the h 1+x , x-diagram
The three cases are exemplarily shown in Fig. 20.27. Starting from the x-line,
the states in the diagram can be fixed by parallel shifting of the corresponding
gradients, see step 2, and the intersection point can be marked. Obviously, the
admixture of water affects the temperature of the final state. Case (3 ) brings
the lowest temperature: the initial state (1) is unsaturated, so it can still absorb a
larger amount of vapour than it currently has. Water is present, but in a solid state.
Consequently, the state of aggregation must be changed from solid to liquid to
vapour. As already known, these phase changes require energy that is taken from
the ambient air (1). Compared to the initial state (1), the temperature therefore
decreases.
The same effect occurs with liquid water, which is added in example (1) → (3 ).
However, the energy requirement is lower compared to (1) → (3 ), since only a
phase change from liquid to vapour has to take place.
Finally, in the example (1) → (3), i.e. with the addition of vapour, no phase change
energy62 but sensible heat is required to bring vapour and air into thermal equilib-
rium. Note that the temperature of the vapour can be adjusted by its pressure, see
steam Tables A.1, A.2, A.3, A.4, A.5, A.6, A.7, A.8, A.9, A.10, A.11 and A.12.

20.4.5 Adiabatic Saturation Temperature

Let us assume unsaturated humid air (1) flows over a surface that is completely
covered with liquid water (2), see Fig. 20.28. Let the initial temperature of the water63
be ϑ2,t=0 , so that the system is not initially in thermodynamic equilibrium. Since the

62 Known as latent heat.


63 The initial temperature of the water ϑ2,t=0 is irrelevant, as the overall system ultimately strives
towards a state of equilibrium. Only the time it takes to reach this state of equilibrium depends on
the initial water temperature.
618 20 Humid Air

Fig. 20.28 Adiabatic saturation temperature—Thermodynamic equilibrium

air is assumed to be unsaturated, as exemplified in Fig. 20.28, it can absorb a further


load of vapour. For this reason, the liquid water begins to vaporise at its surface due
to the low partial pressure of the vapour in the unsaturated air.64 The energy required
for the phase change65 is taken from the surrounding air as well as from the water,
which both cool down to a lower temperature, i.e.

ϑ3 < ϑ1 (20.251)

Ultimately, the system strives towards thermodynamic equilibrium66 :


• Thermal equilibrium
ϑ3 = ϑ2 (20.252)

Due to the thermal equilibrium between water and air, there is no heat exchange at
the water/air interface. On the upper side of the system boundary there shall also
be no heat exchange, see Fig. 20.28, so that the entire process is finally adiabatic.

64 The reason for the mass transfer is the chemical potential between liquid water and unsaturated
air.
65 Evaporation requires a supply of energy.
66 To keep the process in steady state, the evaporating water is compensated by the supply of liquid

water to the water tank.


20.4 Changes of State for Humid Air 619

• Chemical equilibrium
ϕ3 = 1. (20.253)

The air leaves the system boundary in saturated state.


This problem is associated with two possible questions:
(i) What temperature ϑ2 can be reached on the wet surface in steady state?
State (3) is the result of mixing state (1) and state (2), i.e. adding pure liquid
water, see Sect. 20.4.4. Unfortunately, state (2) can not be fixed in a h 1+x , x-
diagram, since liquid water has a water content of x2 → ∞. Anyhow, state (2)
of the liquid water must be located on the ϑ2 -isotherm.
State (1) shall be well known, e.g. by temperature and relative humidity, so that
it can be localised in an h 1+x , x-diagram, see Fig. 20.28. Now, the following
requirements must be met:
1. State (3) must have the same temperature as (2) due to thermodynamic
equilibrium. State (2) is somewhere in the two-phase range.
2. Furthermore, (3) must lie on the saturation line, i.e. ϕ3 = 1.
3. According to the mixing principles, it is further known, see Sect. 20.4.4, that
the direction from (1) to (3) needs to follow

h 1+x
= h w = h liq = cliq · ϑ2 . (20.254)
x
However, in the two-phase region this gradient is identical with the isotherm
ϑ2 , see Eq. 20.187. Thus, this problem can be solved graphically by fixing one
end of a ruler in state (1) and starting to rotate the ruler until one finds a straight
line that includes (1), (2) and (3) and that is congruent with an isotherm in
the two-phase region. Such a line is shown in Fig. 20.28 and fulfils all three
requirements.
In the following, a mathematical solution procedure is presented for the case
that no h 1+x , x-diagram is available. Starting point is state (1) given by ϑ1 and
ϕ1 :
1. Determination of the water content x1 , i.e.

ϕ1 pv (ϑ1 )
x1 = 0.622 . (20.255)
p − ϕ1 pv (ϑ1 )

2. Water balance
ṁ a,1 x1 + ṁ w = ṁ a,1 x3 (20.256)

Thus, the supplied water is

ṁ w = ṁ a,1 (x3 − x1 ) . (20.257)


620 20 Humid Air

With state (3) being saturated, i.e.

pv (ϑ2 )
x3 = x3 = 0.622 . (20.258)
p − pv (ϑ2 )

Thus, one gets




pv (ϑ2 )
ṁ w = ṁ a,1 0.622 − x 1 . (20.259)
p − pv (ϑ2 )

3. Energy balance

ṁ a,1 h 1+x,1 + ṁ w h liq = ṁ a,1 h 1+x,3 (20.260)

Substitution of ṁ w and reducing of ṁ a,1 results in




pv (ϑ2 )
h 1+x,1 + 0.622 − x 1 h liq = h 1+x,3 . (20.261)
p − pv (ϑ2 )

The corresponding specific enthalpies are

h liq = cliq ϑ2 (20.262)

and 
h 1+x,1 = c p,a · ϑ1 + x1 · h v + c p,v · ϑ1 (20.263)

and 
h 1+x,3 = c p,a · ϑ2 + x3 · h v + c p,v · ϑ2 . (20.264)

Equations 20.255 to 20.264 can be solved numerically. The solution is pre-


sented in Fig. 20.29.
On the x-axis the humid air temperature ϑ1 of the flow is shown, the y-axis
shows the adiabatic saturation temperature ϑ2 , that can be achieved in thermo-
dynamic equilibrium. Obviously, with ϕ1 = 100% it is not possible to cool the
surface down, since the air flow (1) is already saturated with vapour, so that
no extra water can evaporate at the liquid surface, i.e. the system has already
reached its equilibrium. Thus, if no phase change takes place, no energy is
consumed, that causes a temperature drop.67 The further the relative humidity
drops, the greater the cooling potential, as the amount of water that evaporates
and then is carried along by the air flow increases. This principle can therefore
be used for cooling purposes.

67As long as no water is transferred into the air flow, the energy balance is not affected, see first
law of thermodynamics, Eq. 20.260.
20.4 Changes of State for Humid Air 621

Fig. 20.29 Adiabatic saturation temperature

(ii) What is the water content of the unsaturated air (1)?


Alternatively, if the temperature of the water ϑ2 is measured and the temperature
of the unsaturated air ϑ1 is known, the principle described in question one can
be used to estimate the water content x1 or the relative humidity ϕ1 of the
unsaturated air. The previously mentioned requirements are still valid. State
(3), however, can easily be fixed in the h 1+x , x-diagram, since

ϑ3 = ϑ2 (20.265)

and
ϕ3 = 1. (20.266)

Now, the ϑ3 -isotherm needs to be extended from the two-phase region to the
one-phase region, cf. Fig. 20.28. Its intersection with the ϑ1 -isotherm in the
one-phase region marks state (1). Finally, x1 can be determined with a h 1+x , x-
diagram. A possible experimental set-up, also known as psychrometer, is shown
in Fig. 20.30. The wet bulb can be realised, for example, by placing the ther-
mometer in a wet cloth.

20.4.6 The h1+x , x-Diagram for Varying Total Pressure

Common h 1+x , x-diagrams can be applied for a fixed total pressure of 1 bar. So the
question is how to apply the diagram for pressures that differ from 1 bar. Suppose a
state (1) with ϑ1 , ϕ1 and p1 = pref = 1 bar is to be represented in a standard h 1+x , x-
diagram which holds for p = pref = 1 bar. Obviously, the relative humidity of a
state
622 20 Humid Air

Fig. 20.30 Psychrometer


(wet-and-dry-bulb
thermometer)

x p
ϕ= · (20.267)
x + 0.622 pv

depends on the total pressure p. For the conversion of p → pref only the pressure
is varied, while temperature ϑ and water content x remain constant.68 The varying
terms are separated on the left-hand side of the equation so that

ϕ x 1
= ·  = const. (20.268)
p x + 0.622 pv
ϕ
Obviously, the ratio p
is constant, i.e. applied for state (1) it is

pref
ϕ1,ref = · ϕ1 (20.269)
p1

Finally, this equation can be applied to plot state (1) with a total pressure of p =
pref = 1 bar in a regular h 1+x , x diagram for p = 1 bar. The only variable that needs
to be converted is the relative humidity, which is a function of the total pressure.
Such a conversion is shown in Fig. 20.31.

Problem 20.3 In the heat exchanger of an air conditioning system, an air flow of
500 kg
h
humid air with a relative humidity of 40% and a temperature of 32.9 ◦ C is

68The maximum partial pressure of the vapour pv remains constant as well, since it exclusively
depends on temperature.
20.4 Changes of State for Humid Air 623

Fig. 20.31 The


h 1+x , x-diagram— p = 1 bar

cooled down. The process runs isobarically at a pressure of 1 bar. The cooling power
is 13.773 MJ
h
.
(a) Sketch the change of state schematically in a h 1+x , x-diagram.
(b) Calculate the water content of the humid air at the inlet of the heat exchanger.
(c) Calculate the mass flow rate of the dry air.
(d) Determine the specific enthalpy, the temperature as well as the water content of
the air when leaving the heat exchanger.
(e) How much liquid water must be removed adiabatically and isobarically to obtain
saturated humid air?
Solution

(a) Sketch of the change of state in a h 1+x , x-diagram:


(b) The water content x1 of state (1) can be determined with a h 1+x , x-diagram, i.e.

x1 = 0.0127. (20.270)

In order to calculate x1 , the following equation needs to be applied

ϕ1 ps (ϑ1 )
x1 = 0.622 = 0.0127. (20.271)
p − ϕ1 ps (ϑ1 )

According to the steam table, the saturated pressure is ps (ϑ1 ) = 0.0501 bar.
(c) The total mass flow rate follows

kg
ṁ = ṁ 1 = ṁ a,1 + ṁ v,1 = 500 . (20.272)
h
624 20 Humid Air

Substitution of the mass of the vapour

ṁ = ṁ 1 = ṁ a,1 + x1 ṁ a,1 (20.273)

and solving for the mass flow rate of the dry air brings

ṁ kg
ṁ a,1 = = 493.83 . (20.274)
1 + x1 h

(d) Application of the first law of thermodynamics69



Q̇ 12 + Pt,12 = ṁ a h 1+x,2 − h 1+x,1 (20.275)

=0

leads to the specific enthalpy of state (2), i.e.

Q̇ 12
h 1+x,2 = h 1+x,1 + . (20.276)
ṁ a

In this equation h 1+x,1 is unknown and can either be determined with the h 1+x , x-
diagram or calculated, i.e.

 kJ
h 1+x,1 = c p,a ϑ1 + x1 h v + c p,v ϑ1 = 65.5 . (20.277)
kg

This leads to the specific enthalpy of state (2) according to Eq. 20.276

Q̇ 12 kJ
h 1+x,2 = h 1+x,1 + = 37.6 . (20.278)
ṁ a kg

The water content x1 , however, is constant, since no water is removed or added,


so that
x2 = x1 = 0.0127. (20.279)

Hence, x2 and h 1+x,2 fix state (2) unambiguously in the h 1+x , x-diagram, cf.
Fig. 20.32, so that its temperature can be determined, i.e.

x2 , h 1+x,2 → ϑ2 = 13.4 ◦ C. (20.280)

Alternatively, temperature ϑ2 can be calculated. The specific enthalpy in state


(2) is

69 There is no technical power Pt,12 to be transferred in the heat exchanger and the change of
kinetic as well as potential energy shall be ignored. Note that a cooling is performed, so that
Q̇ 12 = −13.773 MJh .
20.4 Changes of State for Humid Air 625

Fig. 20.32 Solution to Problem 20.3

kJ  
h 1+x,2 = 37.6 = c p,a ϑ2 + x2 h v + c p,v ϑ2 + x2 − x2 cliq ϑ2 . (20.281)
kg

Solving for the temperature ϑ2

h 1+x,2 − x2 h v
ϑ2 =  (20.282)
c p,a + x2 c p,v + x2 − x2 cliq

with
ps (ϑ2 )
x2 = 0.622 . (20.283)
p − ps (ϑ2 )

Obviously, Eq. 20.282 needs to be solved numerically, since x2 = f (ϑ2 ). The
iteration leads to
ϑ2 = 13.4 ◦ C. (20.284)

(e) The liquid water that needs to be removed follows



ṁ liq = ṁ a x2 − ṁ a x2 = ṁ a x2 − x2 . (20.285)

The water content x2 is

ps (ϑ2 )
x2 = 0.622 = 0.0097. (20.286)
p − ps (ϑ2 )

According to the steam table, the saturated pressure is ps (ϑ2 ) = 0.015 bar. Thus,
the mass flow rate of the liquid water is
626 20 Humid Air

 kg
ṁ liq = ṁ a x2 − x2 = 1.48 . (20.287)
h

3
Problem 20.4 A room needs to be supplied with V̇ = 20000 mh humid air (ϑ4 =
20 ◦ C, ϕ4 = 0.4). This state of air is achieved by adiabatic mixing of heated fresh
air (3) and used air (1). State (1) is characterised by ϑ1 = 25 ◦ C and ϕ1 = 0.6. Fresh
air of state (2) has a temperature of ϑ2 = 0 ◦ C and a relative humidity of ϕ2 = 0.8.
Figure 20.33 shows the technical layout. The total pressure shall be p = 1 bar.
(a) Sketch the changes of state schematically in a h 1+x , x-diagram.
(b) Determine state (3) to which the fresh air (2) must be conditioned.
(c) To what ratio must fresh air and used air be mixed?
(d) What heat flux must be supplied to the fresh air?
Solution
(a) Sketch of the change of state in a h 1+x , x-diagram, see Fig. 20.34.
(b) There are two alternatives to fix state (3):

Fig. 20.33 Technical layout to Problem 20.4

Fig. 20.34 Solution to Problem 20.4


20.4 Changes of State for Humid Air 627

• Alternative 1: Graphical solution


1. Since the change of state from (2) to (3) is pure heating, the water content
remains constant, i.e.
x2 = x3 . (20.288)

Therefore, state (3) must be vertically below state (2).


2. States (1), (3) and (4) must be on a straight line, i.e. mixing straight, due
to the adiabatic mixing, see Sect. 20.4.3. In the task description, however,
states (1) and (4) are given, so that the straight line must be extended to the
left.
3. The graphical construction of 1. and 2. leads to state (3), which is the point
of intersection. It can be fixed in the h 1+x , x-diagram, i.e.

ϑ3 = 17.7 ◦ C (20.289)

and
ϕ3 = 25%. (20.290)

• Alternative 2: Numerical solution


The mass balance of the mixing chamber for air reads as

ṁ a,3 + ṁ a,1 = ṁ a,4 . (20.291)

The following applies accordingly to water

ṁ a,3 x3 + ṁ a,1 x1 = ṁ a,4 x4 . (20.292)

The first law of thermodynamics in steady state for the mixing chamber with
the usual premises is as follows

Ḣ1 + Ḣ3 = Ḣ4 . (20.293)

Introducing the specific enthalpy brings



ṁ a,3 h 1+x,3 + ṁ a,1 h 1+x,1 = ṁ a,3 + ṁ a,1 h 1+x,4 . (20.294)
  
=ṁ a,4

Hence, the specific enthalpy of state (3) is



ṁ a,3 + ṁ a,1 h 1+x,4 − ṁ a,1 h 1+x,1
h 1+x,3 = . (20.295)
ṁ a,3
628 20 Humid Air

For state (1) it is known from the steam table, that

ps (ϑ1 ) = 0.0317 bar. (20.296)

The water content follows


ϕ1 ps (ϑ1 )
x1 = 0.622 = 0.0121. (20.297)
p − ϕ1 ps (ϑ1 )

Its specific enthalpy is

 kJ
h 1+x,1 = c p,a ϑ1 + x1 h v + c p,v ϑ1 = 55.81 . (20.298)
kg

For state (2) it is known from the steam table, that

ps (ϑ2 ) = 0.0061 bar. (20.299)

The water content follows


ϕ2 ps (ϑ2 )
x2 = 0.622 = 0.0031. (20.300)
p − ϕ2 ps (ϑ2 )

We already know, that for state (3) it is

x3 = x2 = 0.0031. (20.301)

For state (4) it is known from the steam table, that

ps (ϑ4 ) = 0.0234 bar. (20.302)

The water content follows


ϕ4 ps (ϑ4 )
x4 = 0.622 = 0.0059. (20.303)
p − ϕ4 ps (ϑ4 )

Its specific enthalpy is

 kJ
h 1+x,4 = c p,a ϑ4 + x4 h v + c p,v ϑ4 = 34.98 . (20.304)
kg

The mass flow rate ṁ a,4 of the dry air can be calculated with


v1+x,4 = (20.305)
ṁ a,4
20.4 Changes of State for Humid Air 629

so that

ṁ a,4 = . (20.306)
v1+x,4

The specific volume of state (4) results in

Ra T4
v1+x,4 = [1 + 1.608x4 ] . (20.307)
p

Hence, the mass flow rate is

p V̇ kg
ṁ a,4 = = 6.539 . (20.308)
Ra T4 [1 + 1.608x4 ] s

Rearranging the mass balance for air, cf. Eq. 20.291, brings

ṁ a,3 = ṁ a,4 − ṁ a,1 . (20.309)

Substitution in the mass balance for water, see Eq. 20.292



ṁ a,4 − ṁ a,1 x3 + ṁ a,1 x1 = ṁ a,4 x4 (20.310)

and solving for


x4 − x3 kg
ṁ a,1 = ṁ a,4 = 2.0479 . (20.311)
x1 − x3 s

The mass flow rate of state (3) follows accordingly

kg
ṁ a,3 = ṁ a,4 − ṁ a,1 = 4.4913 . (20.312)
s
Finally, the specific enthalpy can be calculated, see Eq. 20.295,

ṁ a,3 + ṁ a,1 h 1+x,4 − ṁ a,1 h 1+x,1 kJ
h 1+x,3 = = 25.49 . (20.313)
ṁ a,3 kg

It is further known, that

 kJ
h 1+x,3 = c p,a ϑ3 + x3 h v + c p,v ϑ3 = 25.49 . (20.314)
kg

This equation can be solved for the temperature in state (3), i.e.

h 1+x,3 − x3 h v
ϑ3 = = 17.68 ◦ C. (20.315)
c p,a + x3 c p,v
630 20 Humid Air

The relative humidity is


x3 p
ϕ3 = = 24.15% (20.316)
x3 + 0.622 ps (ϑ3 )

with
ps (ϑ3 ) = 0.0202 bar (20.317)

taken from the steam table.


A comparison of alternative 1 and 2 clearly shows how advantageous the graphic
solution is.
(c) Once again, there are two alternatives:
• Alternative 1: Graphical solution
According to Sect. 20.4.3, the ratio can be fixed with the length of the lever
arms, i.e.
ṁ a,3 14
= = 2.14 (20.318)
ṁ a,1 34

This result has been achieved by measuring the length of the lever arms in a
h 1+x , x-diagram.
• Alternative 2: Numerical solution
The mass flow rates have already been calculated in part (b), i.e.

ṁ a,3 4.4913 kgs


= = 2.193. (20.319)
ṁ a,1 2.0479 kgs

(d) The heat flux results from the application of the first law of thermodynamics
with the usual premises for heating

Q̇ 23 + Pt,23 = ṁ a,2 h 1+x,3 − h 1+x,2 (20.320)

=0

with
 kJ
h 1+x,2 = c p,a ϑ2 + x2 h v + c p,v ϑ2 = 7.64 . (20.321)
kg

Thus, finally the heat flux is



Q̇ 23 = ṁ a,2 h 1+x,3 − h 1+x,2 = 80.19 W. (20.322)
Chapter 21
Steady State Flow Processes

Steady state flow processes have already been discussed in the introduction of the
first law of thermodynamics for open systems, see Sect. 11.3. Figure 21.1 shows
an example for a simple open system with a single inlet and a single outlet. It is
characteristic of open systems in a steady state that the mass in the system remains
constant in time, so that the mass flux into the system must be balanced by the mass
flux out of the system, i.e.
ṁ 1 = ṁ 2 (21.1)

Furthermore, the state inside the system must not vary with respect to time, so that
the first law of thermodynamics obeys

dE
= 0 = Ė in − Ė out . (21.2)
dt
With other words, the energy flux into the system must be equal to the energy flux
out of the system, i.e.
   
1 1
ṁ 1 h 1 + c12 + gz 1 + Q̇ + Pt = ṁ 2 h 2 + c22 + gz 2 (21.3)
2 2

respectively in specific notation

1 2 
q + wt = h 2 − h 1 + c − c12 + g (z 2 − z 1 ) (21.4)
2 2

and in differential notation

δq + δwt = dh + c dc + g dz (21.5)

© Springer Nature Switzerland AG 2022 631


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_21
632 21 Steady State Flow Processes

Fig. 21.1 Steady state flow process

The specific technical work wt can be determined with the partial energy equation,
see Sect. 11.3.4, and follows

2
1 2 
wt = v dp + c2 − c12 + g (z 2 − z 1 ) + ψ12 (21.6)
2
1

Since in a steady state open system no state value may change with time, the entropy
must follow accordingly
dS
= 0 = Ṡin − Ṡout . (21.7)
dt
That is, the entropy flux into the system must be equal to the entropy flux out of the
system, i.e.
Ṡi + Ṡa + ṁ 1 s1 = ṁ 2 s2 (21.8)

respectively in specific notation

si + s a = s 2 − s 1 (21.9)

and in differential notation


δsi + δsa = ds (21.10)

The process values δsi and δsa can be substituted, so that

δψ δq
+ = ds (21.11)
T T

A steady state exergy balance can also be performed for the open system shown, i.e.

dE x
= 0 = ṁ 1 ex,S,1 + Ė x,Q + Pt −  Ė x,V − ṁ 2 ex,S,2 . (21.12)
dt
The exergy flux into the system must be equal to the exergy flux out of the system, i.e.
21 Steady State Flow Processes 633

ṁ 1 ex,S,1 + Ė x,Q + Pt =  Ė x,V + ṁ 2 ex,S,2 (21.13)

with
 Ė x,V = Tenv Ṡi . (21.14)

In specific notation it is

ex,S,1 + ex,Q + wt = ex,V +ex,S,2 (21.15)


 

=Tenv si

In Part I, these equations have been discussed and practised in detail. So far, no
restrictions have been made with regard to the fluid, so that the given correlations are
generally valid. They can be applied for ideal gases as well as for real fluids. In this
chapter, the focus is further on steady state flow processes including supersonic flows.

21.1 Incompressible Flows

Figure 21.2a shows an open system in steady state. Let us assume, that the fluid
is incompressible, e.g. an incompressible liquid as treated in Part I. Incompressible
fluids obey
1
ρ = = const. → v = const. (21.16)
v
Under this premise, the partial energy Eq. 21.6 can be simplified as follows

c22 − c12
wt12 = v ( p2 − p1 ) + ψ12 + + g · (z 2 − z 1 ) . (21.17)
2

Fig. 21.2 Incompressible flow


634 21 Steady State Flow Processes

For a work-insulated flow process, i.e. no technical work is transferred across the
system boundary, it follows:

c22 − c12
wt12 = 0 = v ( p2 − p1 ) + ψ12 + + g · (z 2 − z 1 ) . (21.18)
2
Multiplication with the density leads to

c22 − c12
0 = ( p2 − p1 ) + ρψ12 + ρ + gρ · (z 2 − z 1 ) . (21.19)
2
Rearranging finally brings
ρ 2 ρ
p+ c + gρz − p + c2 + gρz = −ρψ12 ≤ 0 (21.20)
2 2 2 1

Now, let us further assume, the flow is frictionless, i.e.

ψ12 = 0. (21.21)

Under this premise, the partial energy equation can be simplified as follows
ρ 2 ρ
p+ c + gρz − p + c2 + gρz = 0. (21.22)
2 2 2 1

In other words
ρ 2
p+ c + gρz = const. (21.23)
2

This equation is known in fluid dynamics as the Bernoulli equation. It states that in
a frictionless
 flow, where no technical work is transferred, see Fig. 21.2b, the total
pressure1 p + ρ2 c2 + gρz remains constant. If friction occurs, the total pressure
along the flow is reduced by ρψ12 , see Eq. 21.20.

21.2 Adiabatic Flows

In this section adiabatic flows, as shown in Fig. 21.3, are investigated in steady
state.2 The basic principles have already been derived in Part I and are presented

1 The total pressure is composed of the static pressure p to which a person following the flow at its
velocity c would be subjected, the dynamic pressure ρ2 c2 and the hydraulic pressure gρz.
2 The flow cross-section does not have to be constant.
21.2 Adiabatic Flows 635

Fig. 21.3 Adiabatic flow

briefly. Since the flow is adiabatic, there is no exchange of thermal energy with the
environment, i.e.
q12 = 0. (21.24)

Furthermore, a work-insulated flow process without any exchange of technical work


across the system boundary is assumed, so that

wt12 = 0. (21.25)

Under these conditions the first law of thermodynamics for steady state conditions
obeys
c2 − c12
q12 + wt12 = 0 = h 2 − h 1 + 2 + g · (z 2 − z 1 ) . (21.26)
2

With the assumption, that potential energies3 can be ignored, i.e. z 2 = z 1 , it simpli-
fies to
c2 c2
h2 + 2 = h1 + 1 . (21.27)
2 2

Introducing the so-called specific total enthalpy h + = h + c2


2
results in

h+ +
2 = h1 (21.28)

The second law of thermodynamics obeys

ṁs1 + Ṡi + Ṡa = ṁs2 . (21.29)

Since the flow is adiabatic, the exchanged entropy by heat across the system boundary
is zero, i.e. Ṡa = 0. This leads to

ṁ (s1 − s2 ) + Ṡi = 0. (21.30)

3In general, the potential energies can be neglected as long as the vertical distance between inlet
and outlet is not very large.
636 21 Steady State Flow Processes

Dividing by the mass flow rate means

2
δψ
s2 − s1 = si12 = ≥0 (21.31)
T
1

respectively in differential notation

δψ
ds = δsi = ≥0 (21.32)
T

In other words, a decrease in entropy is not possible in an adiabatic flow. The partial
energy equation for the technical work obeys

2
c22 − c12
wt,12 = 0 = v d p + ψ12 + + g · (z 2 − z 1 ) . (21.33)
2
1

In differential notation the partial energy equation yields

δwt = 0 = v d p + δψ12 + c dc + g dz (21.34)

Neglecting the change in potential energy, see above, the partial energy equation
simplifies to
2
c2 − c12
v dp + ψ + 2 =0 (21.35)
2
1

respectively
δwt = 0 = v d p + δψ + c dc (21.36)

Figure 21.4 shows an adiabatic flow process in a h, s-diagram. Let us take state (1)
as a starting point, due to Eq. 21.31 the specific entropy cannot decrease, so the
outlet state must move to the right or vertically. The first law of thermodynamics has
shown, that
c2
h+ = h + = const. (21.37)
2
However, two cases are possible:

• Case 1: Accelerated flow, i.e. dc > 0 → c2 > c1

If the specific enthalpy decreases, the velocity must increase to keep the specific
total enthalpy h + constant. According to the partial energy equation, i.e.

2
c12 − c22
v d p = −ψ12 + <0 (21.38)
 
2

1 ≤0
 
<0
21.2 Adiabatic Flows 637

Fig. 21.4 Adiabatic flow

this case is always accompanied by a pressure loss. The same can be shown with
a differential notation of the partial energy equation, i.e.

δwt = 0 = v d p + δψ + c dc. (21.39)

respectively4
δψ c dc
dp = − − < 0. (21.40)
  v


v
≤0 >0

• Case 2: Decelerated flow, i.e. dc < 0 → c2 < c1

If the specific enthalpy increases, the velocity must decrease to keep the specific
total enthalpy h + constant. For this case, the partial energy equation, i.e.

2
c12 − c22
v d p = −ψ12 + (21.41)
 
2

1 ≤0
 
>0

shows that a pressure increase is possible, when

c12 − c22
ψ12 < (21.42)
2
and a pressure loss, when
c12 − c22
ψ12 > . (21.43)
2

4 The velocity c is always positive, since the direction of the velocity vector has already been taken
into account in the sketches by the flow in or out.
638 21 Steady State Flow Processes

In case that
c12 − c22
ψ12 = (21.44)
2
the pressure remains constant.

21.2.1 Adiabatic Diffusor

Figure 21.5 shows an adiabatic diffuser, which is commonly used to reduce the flow
velocity by widening the cross-section. Thus, the outlet velocity is smaller than the
inlet velocity, i.e.
c2 < c1 . (21.45)

In Sect. 21.2 it has been shown that the first law of thermodynamics in steady state
and neglecting potential energies obeys

c2
h+ = h + = const. (21.46)
2
so that
c12 c2
h1 + = h2 + 2 . (21.47)
2 2
As the flow is decelerated, the specific enthalpy must increase in order to keep the
specific total enthalpy constant, i.e.

h2 > h1. (21.48)

The outlet velocity c2 can be easily calculated and results in



c2 = c12 + 2 (h 1 − h 2 ) (21.49)

Fig. 21.5 Adiabatic diffusor


21.2 Adiabatic Flows 639

The second law of thermodynamics for adiabatic flows, see Sect. 21.2, reads as

δψ
ds = δsi = ≥0 (21.50)
T

so that the specific entropy cannot decrease. Since no technical work is transferred
across the system boundary, the partial energy equation for the technical work is

2
c22 − c12
wt,12 = 0 = v d p + ψ12 + (21.51)
2
1

This leads to a correlation for the change of pressure, i.e.

2
c12 − c22
v d p = −ψ12 + . (21.52)
2
1

However, two cases are possible:


• Case 1: Pressure increase d p > 0, (1) → (2)

c12 − c22
ψ12 < (21.53)
2
• Case 2: Pressure decrease d p < 0, (1) → (3)

c12 − c32
ψ13 > (21.54)
2
These considerations are summarised in a h, s-diagram, see Fig. 21.6. An isentropic
diffusor efficiency is defined as
 
h s h 2s − h 1 0.5 c12 − c2s
2
ηsD = = =   (21.55)
h h2 − h1 0.5 c12 − c22

i.e. the reference is an isentropic change of state (1) → (2s), so that ηsD = 1.

21.2.2 Adiabatic Nozzle

Figure 21.7 shows an adiabatic nozzle used to increase the flow velocity by reducing
the cross-section. Thus, the outlet velocity is greater than the inlet velocity, i.e.
640 21 Steady State Flow Processes

Fig. 21.6 Adiabatic diffusor


in a h, s-diagram

Fig. 21.7 Adiabatic nozzle

Fig. 21.8 Adiabatic nozzle


in a h, s-diagram

c2 > c1 . (21.56)

In Sect. 21.2 it has been shown that the first law of thermodynamics in steady state
and neglecting potential energies obeys, cf. Fig. 21.8,

c2
h+ = h + = const. (21.57)
2
21.2 Adiabatic Flows 641

so that
c12 c2
h1 + = h2 + 2 . (21.58)
2 2
As the flow is accelerated, the specific enthalpy must decrease to keep the specific
total enthalpy constant, i.e.
h2 < h1. (21.59)

The outlet velocity c2 can be easily calculated and equals



c2 = c12 + 2 (h 1 − h 2 ) (21.60)

The second law of thermodynamics for adiabatic flows, see Sect. 21.2, reads as

δψ
ds = δsi = ≥0 (21.61)
T

so that the specific entropy needs to increase. Since no technical work is transferred
across the system boundary, the partial energy equation for the technical work is

2
c22 − c12
wt,12 = 0 = v d p + ψ12 + (21.62)
2
1

This results in a correlation for the pressure change, i.e.

2
c12 − c22
v d p = −ψ12 + < 0. (21.63)
 
2

1 ≤0
 
<0

With an adiabatic nozzle, the pressure must therefore decrease. These considerations
are summarised in a h, s-diagram, see Fig. 21.6. An isentropic diffusor efficiency is
defined as  
h h2 − h1 0.5 c12 − c22
ηsN = = =   (21.64)
h s h 2s − h 1 0.5 c12 − c2s
2

i.e. the reference is an isentropic change of state (1) → (2s), so that ηsN = 1.

Problem 21.1 A mass flux of 10 kgs vapour with an initial state of p1 = 22 bar and
ϑ1 = 500 ◦ C shall expand polytropically with a constant exponent of n = 1.2 in
adiabatic nozzle to a pressure of p2 = 15 bar. The inlet velocity of the vapour is
c1 = 10 ms . Changes of the potential energy can be ignored. Please calculate:
642 21 Steady State Flow Processes

(a) the specific volume v2 and the temperature ϑ2 of the vapour in state (2),
(b) the outlet velocity c2 of the vapour,
2
(c) the specific pressure work y12 = v d p and the specific dissipation ψ12 ,
1
(d) the thermodynamic mean temperature,
(e) the cross-sectional area of inlet and outlet.

Solution

(a) According to the steam table, state (1) is superheated5 steam with

m3
v1 = 0.1595 . (21.65)
kg

For a polytropic change of state it is6

pvn = const. (21.66)

This leads to   n1
p1 m3
v2 = v1 = 0.2194 . (21.67)
p2 kg

According to the steam table one gets

ϑ2 = f (v2 , p2 ) = 450.72 ◦ C. (21.68)

Note that   n−1


p2 n
T2 = T1 (21.69)
p1

cannot be applied because this equation is derived by applying the ideal gas law.
(b) The first law of thermodynamics for the nozzle reads as
⎡ ⎤

1 2 
Q̇ 12 + Pt,12 = 0 = ṁ ⎣h 2 − h 1 + c − c12 + g (z 2 − z 1 )⎦ . (21.70)
2 2  

=0

Thus, it follows
c2 = 2 (h 1 − h 2 ) + c12 . (21.71)

5 This is due to ϑ1 > ϑ  ( p1 ).


6 No matter what fluid is used, cf. Sect. 13.6.
21.2 Adiabatic Flows 643

The specific enthalpies follow from the steam table, i.e.

kJ
h 1 = h (22 bar, 500 ◦ C) = 3465.9 (21.72)
kg

and
kJ
h 2 = h (15 bar, 450.72 ◦ C) = 3366.2 . (21.73)
kg

Finally, the outlet velocity is7



m
c2 = 2 (h 1 − h 2 ) + c12 = 446.57 . (21.74)
s
(c) Part (a) has shown, that for a polytropic change of state it always follows
  n1
p1 m3
v = v1 = 0.2194 . (21.75)
p kg

The pressure work can be calculated accordingly, i.e.

2   n−1 
n p2 n kJ
y12 = v d p = v1 p1 − 1 = −130.16 . (21.76)
n−1 p1 kg
1

The technical work then is

2
1 2 
wt,12 = v d p + ψ12 + c − c12 = 0. (21.77)
2 2
1

Note that no technical work is transferred8 across the system boundary of the
nozzle. Thus, the dissipation is

2 2
1 2  kJ
ψ12 = c1 − c22 − v d p = h2 − h1 − v d p = 30.49 . (21.78)
2 kg
1 1

(d) The second law of thermodynamics obeys

ṁs1 + Ṡi + Ṡa = ṁs2 . (21.79)

7 J
Note the units. The specific enthalpies should be converted to kg .
8 A nozzle is a work-insulated, i.e. passive, system.
644 21 Steady State Flow Processes

Since the nozzle is adiabatic it simplifies to

Ṡi = ṁ (s2 − s1 ) . (21.80)

In specific notation it results in

2 2
δψ 1 ψ12
s i = s2 − s1 = = δψ = . (21.81)
T Tm Tm
1 1

This results in the following for the thermodynamic mean temperature

ψ12
Tm = . (21.82)
s2 − s1

The specific entropies follow from the steam table, i.e.

kJ
s1 = s (22 bar, 500 ◦ C) = 7.3873 (21.83)
kg K

and
kJ
s2 = s (15 bar, 450.72 ◦ C) = 7.4281 . (21.84)
kg K

Finally, the thermodynamic mean temperature is

ψ12
Tm = = 748.36 K (21.85)
s2 − s1

respectively
ϑm = 475.21 ◦ C. (21.86)

(e) The mass flux needs to be constant in steady state, so that

Ac
ṁ = ρ V̇ = ρ Ac = = const. (21.87)
v
This leads to
ṁv1
A1 = = 0.1595 m2 (21.88)
c1

and
ṁv2
A2 = = 49.13 cm2 . (21.89)
c2
21.3 Velocity of Sound 645

21.3 Velocity of Sound

For the following sections it is required to have knowledge of the velocity of sound a.
The velocity of sound describes the speed at which a pressure pulse travels through
a fluid. To derive this velocity, a fluid is assumed to move unimpeded at a velocity
c, see Fig. 21.9. At a certain point in time, a pressure pulse is triggered, causing a
wave that moves with a velocity a. This situation is shown in Fig. 21.9a with a fixed
coordinate system. To the left of the impulse, the fluid is undisturbed and enters the
coordinate system with a pressure p, a temperature T , a density ρ and a velocity c.
However, each of these state values has changed when leaving the control volume,
i.e. the pressure is p + d p, the temperature is T + dT , the density ρ + dρ and the
velocity is c + dc.
In a next step, the coordinate system starts to move, cf. Fig. 21.9b. The velocity of
the moving coordinate system can be adjusted9 so that the fluid enters the coordinate
system with velocity a on the left, i.e. according to the sketch, the movement velocity
of the coordinate system must be a − c. This results in the inlet and outlet velocities
of Fig. 21.9b:
• Left of the control volume:

c +

(a − c) = 

a (21.90)
 

a) Velocity of co-system b)

Fig. 21.9 Velocity of sound a

9 Imagine a pressure pulse moving from right to left with a velocity a = 100 km h and is subjected to
a headwind of c = 20 km h . Thus, the fluid hits the pulse front with a velocity of 120 km
h . However,
the velocity of the fluid behind the impulse shall be slightly higher, e.g. c + dc = 22 km h , so that
the velocity increases by dc = 2 km h . If the air is to hit the pulse front with exactly the velocity
a = 100 kmh , the coordinate system must move to the left with a velocity of a − c = 80 km
h . In this
case, the velocity of the air leaving the coordinate system on the right side is a + dc = 102 km h .
Note that any coordinate system can be used for balancing. The solution must never depend on the
choice of coordinate system.
646 21 Steady State Flow Processes

• Right of the control volume:

 dc
+
c + (a − c) = a +
 dc
(21.91)
 

a) Velocity of co-system b)

The equation of continuity, i.e. principle of mass conservation, for the moving coor-
dinate system, Fig. 21.9b, reads as

Aρa = A (ρ + dρ) (a + dc) . (21.92)

Ignoring the second order non-linear term dρ dc, yields

ρ dc = −a dρ. (21.93)

The conservation of momentum for the moving coordinate system is, see [37],
  
d I
= ρ c (
c · n) dA = F = − n p dA. (21.94)
dt
A A

In this case the problem is one-dimensional, so that

A (ρ + dρ) (a + dc) (a + dc) − Aρaa = p A − ( p + d p) A. (21.95)


 

Aρa

In combination with the mass conservation it follows

Aρa (a + dc) − Aρaa = p A − ( p + d p) A. (21.96)

Rearranging brings
dp
ρ dc = − . (21.97)
a
A combination of equation of continuity and conservation of momentum results in

dp
a2 = . (21.98)

The momentum equation has been applied without friction, i.e. the flow is reversible.
If it is further assumed that the pulse is moving so fast that there is no time for heat
exchange, the flow is also adiabatic. It therefore follows that the process is isentropic.
Finally, the velocity of sound a is
 
dp
a= . (21.99)
dρ s
21.3 Velocity of Sound 647

For an isentropic change of state of an ideal gas it is known that


p
pvκ = = const. → p = Cρ κ . (21.100)
ρκ

The derivation for an isentropic change of state is

dp
= Cκρ κ−1 . (21.101)

The constant C can be substituted by, cf. Eq. 21.100,


p
C= . (21.102)
ρκ

This finally leads to


dp p p
= κ κρ κ−1 = κ = κ pv. (21.103)
dρ ρ ρ

Hence, the velocity of sound for an ideal gas follows

√ √
a= κ pv = κ RT (21.104)

The dimensionless Mach-number Ma is defined by the ratio

c
Ma = (21.105)
a

The same correlation as in Eq. 21.103 can be derived with the second law of ther-
modynamics. In differential notation it obeys for an ideal gas

dT dv
ds = δsi + δsa = 0 = cv +R . (21.106)
T v
With the thermal equation of state for an ideal gas
pv p v
T = → dT = dv + d p (21.107)
R R R
one gets with c p = R + cv
dv dp
0 = cp + cv . (21.108)
v p

It is further known, that


1
v = ρ −1 → dv = − dρ. (21.109)
ρ2
648 21 Steady State Flow Processes

For an isentropic pressure pulse, the ratio of pressure change to density change must
therefore be as follows
dp p
= κ = κ pv. (21.110)
dρ ρ

The pressure pulse in Fig. 21.9b represents an imperfection that is usually associated
with the generation of entropy, see Part I. However, it is assumed that this imperfection
is moving so fast that the system is not able to respond to that imperfection. This is
the reason why no entropy generation has been assumed and why conservation of
momentum has been applied without friction.

21.4 Fanno Correlation

Let us assume an adiabatic but frictional flow in a tube with constant cross-section A,
see Fig. 21.10. Such a flow is denoted Fanno10 flow. The continuity equation states
that the mass flux is constant in the steady state, i.e.

Ac
ṁ = ρ Ac = = const. (21.111)
v
In other words, the mass flux density is also constant because the cross-section does
not vary.
c ṁ
= = ṁ  = const. (21.112)
v A

The first law of thermodynamics for steady state conditions obeys11

1 2 
q + wt = 0 = h 2 − h 1 + c2 − c12 (21.113)
2

Fig. 21.10 Adiabatic,


frictional tube flow
(A = const.)

10 Gino Girolamo Fanno (18 November 1882 in Conegliano, 23 March 1962 in Pegli).
11 The potential energies are neglected. Also, no technical work occurs across the system boundary.
2.4 Fanno Correlation 649

with the so-called specific total enthalpy h + it reads as

c2
h+ = h + = const. = h +
1. (21.114)
2

The specific total enthalpy of the inlet shall be given12 and is h +


1 . However, the
combination of the first law of thermodynamics and the conservation of mass leads to

1
h + v2 ṁ 2 = const. = h +
1 (21.115)
2

According to the second law of thermodynamics for adiabatic flow it is

δψ
ds = δsi = ≥0 (21.116)
T

In this case, the specific entropy must therefore increase due to friction. A reduction
of entropy is not possible because the flow is supposed to be adiabatic. Furthermore,
the partial energy equation under these conditions is

2
1 2 
wt = 0 = v dp + c2 − c12 + ψ12 . (21.117)
2
1

Unfortunately, this equation cannot be applied further because the friction is unknown.

The idea now is to represent the change of state of an ideal gas in a h, s-diagram:
1. Inlet state (1) is fixed by two independent state values, e.g. T1 and p1 . Thus, the
specific volume is
RT1
v1 = . (21.118)
p1

The specific enthalpy13 is

h 1 = h ref + c p (T1 − Tref ) . (21.119)

The specific entropy obeys

T1 p1
s1 = sref + c p ln − R ln . (21.120)
Tref pref

Hence, state (1) can be fixed unambiguously in a h, s-diagram.

12 For instance by its internal state p1 , T1 and its velocity c1 .


13 Tref , pref , h ref and sref represent a arbitrary reference state.
650 21 Steady State Flow Processes

2. Next, the velocity c1 is chosen so that the mass flux density is defined as

c1 ṁ 
=  = ṁ (a) = const. (21.121)
v1 A 1

3. Now the specific volume v is varied continuously while the mass flux density is
kept constant. Thus a new velocity can be calculated according to
c1
c (v) = v = vṁ (a) . (21.122)
v1

The updated specific enthalpy then follows, see Eq. 21.114,

1 2 c12 c2 ṁ 2
(a)  2 
h (v) = h +
1 − c = h1 + − = h1 + v1 − v2 . (21.123)
2 2 2 2
The temperature is
h (v) − h ref
T (v) = Tref + (21.124)
cp

and the pressure follows


RT
p (v) = . (21.125)
v
Finally, the new specific entropy is

T (v) p (v)
s (v) = sref + c p ln − R ln . (21.126)
Tref pref

4. The new state (h, s) can be fixed in the h, s-diagram for varying specific volumes.
5. The previous steps 2–4 can now be repeated for a different velocity c1 , i.e. for a
different mass flux density ṁ (b) .
Following these steps, the corresponding curves are plotted in an h, s-diagram. The
curves are called Fanno-curves, see Figs. 21.11 and 21.13. However, a distinction is
made with regard to the inlet condition (1):
• Subsonic inlet, i.e. Ma1 < 1

Under these conditions the given algorithm, see steps 1–5, leads to the curves as
shown in Fig. 21.11. Starting at the inlet state (1), the specific entropy starts rising.
This characteristic makes sense because the friction leads to an increase in specific
entropy when passing through the pipe. In addition, the pressure decreases, as
shown by the isobars in the h, s-diagram, cf. Fig. 21.11. Figure 21.12 additionally
proves that both the velocity and the specific volume start to increase.14 At a point

14 Compare with Problem 12.5.


2.4 Fanno Correlation 651

Fig. 21.11 Subsonic Fanno-curves, ideal gas with κ = const.

Fig. 21.12 Subsonic Fanno-curves for one mass flux density, ideal gas with κ = const.

(A) the flow finally reaches the velocity of sound, i.e. Ma = 1. However, any
further acceleration would then lead to a decrease in specific entropy. This would
violate the second law of thermodynamics, so the dashed lines are possible from
a mathematical point of view, but not from a thermodynamic point of view.

Theorem 21.1 Due to friction the velocity in an adiabatic flow with constant cross-
section and subsonic inlet conditions rises. The maximum velocity that can be
achieved is the speed of sound, i.e. Ma = 1.
652 21 Steady State Flow Processes

The outlet pressure can not be reduced under its critical pressure pA in state (A), no
matter what ambient pressure rules. Thus, to further reduce the pressure, the mass flux
density needs to be lowered and the Fanno-curve shifts to the right in a h, s-diagram
until the desired outlet pressure occurs. Normal shocks, as described in Sect. 21.6,
can not occur, since the flow is subsonic.
So far no statement has been made regarding the length of the tube. However,
for a subsonic flow, there is a critical length, where Ma = 1 is achieved. In case the
tube is now extended beyond the critical length, at which the specific entropy can no
further rise, the new outlet velocity still is Ma = 1. However, this comes along with
a reduction of the mass flow rate, i.e. also with a reduction of the initial velocity. The
flow is blocked by frictional effects (“choking”), the Fanno-curve is shifted to the
right in a h, s-diagram. The frictional blocking leads to a decrease of the effective
cross-section. Thus, a supersonic state can not be achieved, as this would require a
cross-section expansion, see Sect. 21.7.2.

Example 21.1 Air flows through an adiabatic, horizontal channel. The channel
cross-section remains constant, air can be considered an ideal gas (c p = 1004 kgJK ,
R = 287 kgJK → κ = 1.4003). The inlet state (1) shall be:
• T1 = 300 K
• p1 = 5 bar
• Ma1 = 0.05
The specific entropy generation due to dissipation is si,12 = 200 kgJK . Goal is to
determine the Ma-number in state (2). First, the reference state for the caloric state
values is fixed. This choice is basically arbitrary, but may not be changed during the
change of state:
• h 0 (T0 = 10 K) = 0 kgJ
• s0 (T0 = 10 K, p0 = 1 bar) = 0 kgJK
In a next step, the missing state values for state (1) are calculated:
 m
c1 = Ma1 a1 = Ma1 κ RT1 = 17.3612 (21.127)
s

RT1 m3
v1 = = 0.1722 (21.128)
p1 kg

J
h 1 = c p (T1 − T0 ) + h 0 = 2.9116 × 105 (21.129)
kg

T1 p1 J
s1 = s0 + c p ln − R ln = 2.9529 × 103 (21.130)
T0 p0 kg K

To determine state (2) the following set of equations must be solved:


2.4 Fanno Correlation 653

• First law of thermodynamics

1 J 1
h 1 + c12 = 2.9131 × 105 = h 2 + c22 (21.131)
2 kg 2

• Second law of thermodynamics

J T2 p2
s2 = s1 + si,12 = 3.1529 × 103 = s0 + c p ln − R ln (21.132)
kg K T0 p0

• Caloric equation of state


h 2 = c p (T2 − T0 ) + h 0 (21.133)

• Thermal equation of state


p2 v2 = RT2 (21.134)

• Equation of continuity

c1 c2 kg
ṁ 1 = = = 100.8198 (21.135)
v1 v2 s m2

The system of equations is determined: there are five equations for five unknowns.
The numerical solution is:
• T2 = 299.5406 K
• p2 = 2.4774 bar
• c2 = 34.9853 ms
3
• v2 = 0.3470 mkg
• h 2 = 2.9070 × 105 J
kg

The Ma-number for state (2) therefore is


c2 c2
Ma2 = =√ = 0.1008. (21.136)
a2 κ RT2

Obviously, the flow has been accelerated. The change of state is polytropic with an
exponent of
ln pp1
n = v22 = 1.0022. (21.137)
ln v1

Consequently, the specific pressure work is


 
RT1 T2 J
y12 = n − 1 = −6.0415 × 104 . (21.138)
n−1 T1 kg
654 21 Steady State Flow Processes

The application of the partial energy equation finally leads to the specific dissipation:

1 2  J
ψ12 = −y12 − c2 − c12 = 5.9954 × 104 . (21.139)
2 kg

• Supersonic inlet, i.e. Ma1 > 1

Under these conditions the given algorithm, see steps 1–5, leads to the curves as
shown in Fig. 21.13. Starting at the inlet state (1), the specific entropy starts rising.
This characteristic makes sense, as friction leads to an increase in specific entropy
when flowing through the pipe. Furthermore, the pressure increases as proven by
the isobars in the h, s-diagram, see Fig. 21.13. According to the partial energy
equation, see Eq. 21.117,

2
1 2 
v dp = − c − c12 − ψ12 > 0 (21.140)
2 2
1

the change of pressure does not purely depend on the dissipation but also on the
change of velocity, i.e. the flow needs to decelerate, i.e. c2 < c1 :

c2 < c12 − 2ψ12 . (21.141)

Figure 21.14 additionally proves that the velocity as well as the specific volume
decrease. At a point (A) the flow finally has decelerated to the speed of sound,
i.e. Ma = 1. However, any further deceleration would then lead to a decrease of

Fig. 21.13 Supersonic Fanno-curves, ideal gas with κ = const.


2.4 Fanno Correlation 655

Fig. 21.14 Supersonic Fanno-curves for one mass flux density, ideal gas with κ = const.

entropy. This would violate the second law of thermodynamics, so the dashed lines
are possible from a mathematical point of view, but not from a thermodynamic
point of view.

Theorem 21.2 Due to friction the velocity in an adiabatic flow with constant cross-
section and supersonic inlet conditions drops. The minimum velocity that can be
achieved is the speed of sound, i.e. Ma = 1.

From these considerations it follows that the outlet pressure cannot exceed the critical
pressure in state (A). Thus, to further increase the pressure, the mass flux density
needs to be increased and the Fanno-curve shifts to the right in a h, s-diagram until
the desired outlet pressure occurs. Depending on the ambient pressure a normal
shock can occur, see Sect. 21.6, so that the supersonic flow compresses in a unsteady
change of state to a higher pressure level.
In case the flow is supersonic, there is Ma = 1 after a critical length due to fric-
tional effects. Let us now extend the tube beyond this critical length: The flow now
adjusts with a normal shock, so that the flow becomes subsonic, cf. Sect. 21.6. Start-
ing from that new subsonic state the flow now is accelerated, so that it again reaches
Ma = 1 at its end, see also [38].

Example 21.2 Air flows through an adiabatic, horizontal channel. The channel
cross-section remains constant, air can be considered an ideal gas (c p = 1004 kgJK ,
R = 287 kgJK → κ = 1.4003). The inlet state (1) shall be:

• T1 = 500 K
• p1 = 5 bar
• Ma1 = 1.8
656 21 Steady State Flow Processes

The specific entropy generation due to dissipation is si,12 = 100 kgJK . Goal is to
determine the Ma-number in state (2). First, the reference state for the caloric state
values is fixed. This choice is basically arbitrary, but may not be changed during the
change of state:
• h 0 (T0 = 10 K) = 0 kgJ
• s0 (T0 = 10 K, p0 = 1 bar) = 0 kgJK
In a next step, the missing state values for state (1) are calculated:
 m
c1 = Ma1 a1 = Ma1 κ RT1 = 806.874 (21.142)
s
RT1 m3
v1 = = 0.2870 (21.143)
p1 kg
J
h 1 = c p (T1 − T0 ) + h 0 = 4.91960 × 105 (21.144)
kg
T1 p1 J
s1 = s0 + c p ln − R ln = 3.4658 × 103 (21.145)
T0 p0 kg K

To determine state (2) the following set of equations must be solved:


• First law of thermodynamics

1 J 1
h 1 + c12 = 8.1748 × 105 = h 2 + c22 (21.146)
2 kg 2

• Second law of thermodynamics

J T2 p2
s2 = s1 + si,12 = 3.5658 × 103 = s0 + c p ln − R ln (21.147)
kg K T0 p0

• Caloric equation of state


h 2 = c p (T2 − T0 ) + h 0 (21.148)

• Thermal equation of state


p2 v2 = RT2 (21.149)

• Equation of continuity

c1 c2 kg
ṁ 1 = = = 2.8114 × 103 (21.150)
v1 v2 s m2

The system of equations is determined: there are five equations for five unknowns.
The numerical solution is:
• T2 = 653.7839 K
• p2 = 9.0172 bar
2.4 Fanno Correlation 657

• c2 = 585.0194 ms
3
• v2 = 0.2081 mkg
• h 2 = 6.4636 × 105 J
kg

The Ma-number for state (2) therefore is


c2 c2
Ma2 = =√ = 1.1413. (21.151)
a2 κ RT2

Obviously, the flow has been decelerated. The change of state is polytropic with an
exponent of
ln pp21
n = v2 = 1.8341. (21.152)
ln v1

Consequently, the specific pressure work is


 
RT1 T2 J
y12 =n − 1 = 9.7053 × 104 . (21.153)
n−1 T1 kg

The application of the partial energy equation finally leads to the specific dissipation:

1 2  J
ψ12 = −y12 − c2 − c12 = 5.7346 × 104 . (21.154)
2 kg

Both cases have shown that the velocity at point (A) is the velocity of sound, which
is now to be proven. However, Figs. 21.11 and 21.13 indicate that the Fanno-curves
have a vertical tangent at point (A), i.e. dsA = 0. At that point the entropy can not
further rise. Hence, the fundamental equation of thermodynamics, see Sect. 12.3.2,
in point (A) reads as
T ds = dh − v d p = 0. (21.155)

In differential notation the first law of thermodynamics obeys

dh + c dc = 0. (21.156)

The equation of continuity cρ = const. leads to

d (cρ) = ρ dc + c dρ = 0. (21.157)

A combination of these three equation leads to the velocity in point (A)



dp
c= =a (21.158)

658 21 Steady State Flow Processes

The velocity at point (A) must therefore be the velocity of sound. Since no restrictions
have been placed on the inlet conditions, it applies to both subsonic and supersonic
inlet states (1).
Following the first law of thermodynamics, state (A) obeys

1 1 1 1
h A + cA2 = h 1 + c12 → h 1 − h A + c12 = cA2 . (21.159)
2 2 2 2
With the caloric equation of state it results in

1 1
c p (T1 − TA ) + c12 = cA2 . (21.160)
2 2
The velocity in (A) is the velocity of sound, i.e.

cA = κ RTA (21.161)

A combination of both equations leads to

c p T1 + 21 c12 c p T1 + 21 ṁ 2 v12 


TA = = → cA = κ RTA (21.162)
1
2
κR + cp 1
2
κR + cp

The conservation of mass yields



v1 v1  κ RTA
vA = cA = κ RTA = (21.163)
c1 c1 ṁ 

The pressure follows the thermal equation of state

RTA
pA = (21.164)
vA

The caloric state values can then be calculated

TA pA
sA = sref + c p ln − R ln (21.165)
Tref pref

and
h A = h ref + c p (TA − Tref ) (21.166)

Each Fanno-curve is a function of an initial state (1) and a mass flux density ṁ  , i.e.
its velocity c1 .
21.5 Rayleigh Correlation 659

21.5 Rayleigh Correlation

Let us now assume a non-adiabatic but frictionless tube flow with constant cross-
section A, see Fig. 21.15. The continuity equation states that the mass flux is constant
in steady state, i.e.
Ac
ṁ = ρ Ac = = const. (21.167)
v
In other words, the mass flux density is also constant because the cross-section does
not change, i.e.
c ṁ
= = ṁ  = const. (21.168)
v A

The first law of thermodynamics for steady state conditions obeys

1 2 
q + wt = h 2 − h 1 + c − c12 . (21.169)

2 2
=0

However, this is not helpful for a non-adiabatic flow because the specific heat q
is often not known. According to the second law of thermodynamics for diabatic,
reversible flows, the following applies

δq
ds = δsa = ≶0 (21.170)
T

In this case, entropy can therefore increase due to heating or decrease due to cooling.15
The partial energy equation under these conditions is

2
1 2 
wt = 0 = v dp + c − c12 + ψ12 . (21.171)
2 2 

1 =0

Fig. 21.15 Non-adiabatic,


frictionless tube flow
(A = const.)

15 Sure, it can remain constant, but then there would be no driver for any changes of the flow.
660 21 Steady State Flow Processes

In contrast to the derivation of the Fanno correlation, this equation can now be applied
because the flow is assumed to be frictionless. In differential notation it reads as

v d p + c dc = 0. (21.172)

Division by the specific volume yields


c
dp + dc = d p + ṁ  dc = 0. (21.173)
v
The integration leads to

p2 − p1 + ṁ  (c2 − c1 ) = 0. (21.174)

Consequently, the following total pressure needs to be constant, i.e.

p2 + ṁ 2 v2 = p1 + ṁ 2 v1 = const. (21.175)

Since the cross-section is constant this equation can be interpreted as conservation


of momentum:
  
d I
= ρ c (
c · n) dA = F = − n p dA. (21.176)
dt
A A

This leads to
ρ2 c22 A − ρ1 c12 A = p1 A − p2 A. (21.177)

However, Eqs. 21.175 and 21.177 are identical.

As with the Fanno curves, the task now is to represent the change of state of an ideal
gas in a h, s-diagram:
1. Inlet state (1) is fixed by two independent state values, e.g. T1 and p1 . Thus, the
specific volume is
RT1
v1 = . (21.178)
p1

The specific enthalpy16 is

h 1 = h ref + c p (T1 − Tref ) . (21.179)

16 Tref , pref , h ref and sref represent an arbitrary reference state.


21.5 Rayleigh Correlation 661

The specific entropy obeys

T1 p1
s1 = sref + c p ln − R ln . (21.180)
Tref pref

Hence, state (1) can be fixed unambiguously in a h, s-diagram.


2. Next, the velocity c1 is chosen so that the mass flux density is defined as

c1 ṁ 
=  = ṁ (a) = const. (21.181)
v1 A 1

3. Now, the velocity c is varied continuously while keeping the mass flux density
ṁ (a) constant. Thus, a new specific volume can be calculated according to

v1 c
v (c) = c =  . (21.182)
c1 ṁ (a)

The updated pressure then follows, see Eq. 21.175,

p (c) = p1 − ṁ 2
(a) (v (c) − v1 ) . (21.183)

The temperature is
pv (c)
T (c) = (21.184)
R
and the specific enthalpy follows

h (c) = h ref + c p (T (c) − Tref ) . (21.185)

Finally, the new specific entropy is

T (c) p (c)
s (c) = sref + c p ln − R ln . (21.186)
Tref pref

4. The new state (h, s) can be fixed in the h, s-diagram for varying velocities.
5. The previous steps 2–4 can now be repeated for a different velocity c1 , i.e. for a
different mass flux density ṁ (b) .
Following these steps, the corresponding curves are plotted in a h, s-diagram. The
curves are denoted Rayleigh-curves, see Figs. 21.16 and 21.20. However, a distinction
is made with regard to the inlet condition (1):

• Subsonic inlet, i.e. Ma1 < 1

Under these conditions the given algorithm, see steps 1–5, leads to the curves as
shown in Fig. 21.16. Starting at the inlet state (1), the specific entropy rises when
heat is supplied. According to the isobars in Fig. 21.16 the pressure decreases
662 21 Steady State Flow Processes

Fig. 21.16 Subsonic Rayleigh-curves for heating, ideal gas with κ = const.

Fig. 21.17 Subsonic Rayleigh-curves for one mass flux density and heating, ideal gas with κ =
const.

under these conditions. Figure 21.17 additionally proves, that the velocity rises
while heat is supplied. At a point (A) the flow finally has reached the speed of
sound, i.e. Ma = 1. However, any further acceleration would then lead to a decrease
of entropy, i.e. it is impossible for a heated tube.

Theorem 21.3 The maximum heat supply is limited. In order to increase the heating,
the mass flux density needs to be decreased. This results in a reduced pressure.
Heating can lead to a thermal blocking of the subsonic flow, i.e. the flow causes
an unsteady effect similar to a reduction of the cross section, so that its initial state
21.5 Rayleigh Correlation 663

Fig. 21.18 Rayleigh correlation (subsonic), p, v-diagram

respectively the mass flow rate is adjusted, see [38]. It is not possible to accelerate
the fluid thermally beyond state (A), see [39]. Interestingly, close to the critical
point (A)17 the maximum specific enthalpy is reached before the maximum specific
entropy is reached, see Fig. 21.16. In other words, once h max is reached, a further
heating leads to a temperature reduction due to the large expansion of the fluid. The
expansion comes along with a fluid acceleration, see Eq. 21.168. Hence, the first law
of thermodynamics in differential notation yields

δq = dh + c dc. (21.187)

The closer one gets to state (A) the larger c dc gets. A combination the first law of
thermodynamics with the partial energy equation, cf. Eq. 21.172, results in

δq = dh − v d p (21.188)

respectively
δq = du + v d p + p dv − v d p = cv dT + p dv. (21.189)

The Rayleigh-correlation can additionally be illustrated in a p, v-diagram, see


Fig. 21.18. Following Eq. 21.175, the gradient of the Rayleigh-correlation in a p, v-
diagram is constant, i.e.

17 Thus, at high subsonic velocities.


664 21 Steady State Flow Processes

dp
= −ṁ 2 = const. (21.190)
dv
Although the focus so far has been on heating, the p, v-diagram, see Fig. 21.18, also
contains information for the cooling case. According to the first law of thermody-
namics, it follows that cooling leads to a decrease in specific enthalpy, i.e.
Theorem 21.4 A subsonic Rayleigh flow is accelerated by heating and decelerated
by cooling.
Obviously, the polytropic exponent cannot be constant in this case. Its variation over
the specific entropy is shown in Fig. 21.19a. Once, state (A) is achieved, it is n = κ
since no heat can be supplied any more and the flow is frictionless, i.e. isentropic
conditions occur.
• Supersonic inlet, i.e. Ma1 > 1

Under these conditions the given algorithm, see steps 1–5, leads to the curves as
shown in Fig. 21.20. Starting at the inlet state (1), the specific entropy rises as
entropy is carried into the system by the supplied heat. Furthermore, the pressure
increases. According to the partial energy equation

2
1 2 
v dp = − c2 − c12 > 0 (21.191)
2
1

the rise of pressure is because the flow decelerates, i.e. c2 < c1 . Figure 21.21
additionally proves, that the velocity decreases. At a point (A) the flow finally has

Fig. 21.19 Polytropic exponent—a subsonic, b supersonic


21.5 Rayleigh Correlation 665

Fig. 21.20 Supersonic Rayleigh-curves for heating, ideal gas with κ = const.

Fig. 21.21 Supersonic Rayleigh-curves for one mass flux density and heating, ideal gas with
κ = const.

decelerated to the speed of sound, i.e. Ma = 1. However, a further deceleration


would require a decrease in entropy, which is not possible in a heated tube.
Theorem 21.5 The maximum heat supply is limited. In order to increase the heating,
the mass flux density needs to be increased. This leads to an enlarged pressure.
When Ma = 1 is reached and heat is further supplied, a normal shock takes place,
see Sect. 21.6. By this normal shock the flow is then subsonic and starts to accelerate
again to Ma = 1, see [38]. In case heating is further increased the normal shock moves
666 21 Steady State Flow Processes

Fig. 21.22 Rayleigh correlation (supersonic), p, v-diagram

upstream until the entire tube has subsonic velocity. The Rayleigh-correlation can
additionally be illustrated in a p, v-diagram, see Fig. 21.22. Same as for the subsonic
flow, following Eq. 21.175, the slope of the Rayleigh-correlation in a p, v-diagram
is constant, i.e.
dp
= −ṁ 2 = const. (21.192)
dv
Though the focus has been on heating so far, the p, v-diagram 21.22 also contains
information regarding the cooling case. According to the first law of thermodynamics,
it follows that cooling leads to a decrease in specific enthalpy, i.e.

Theorem 21.6 A supersonic Rayleigh flow is accelerated by cooling and deceler-


ated by heating.

Obviously, the polytropic exponent cannot be constant in this case. Its variation over
the specific entropy is shown in Fig. 21.19b.

Both cases have shown that the velocity at point (A) is the velocity of sound, which is
now to be proven. However, Figs. 21.16 and 21.20 indicate that the Rayleigh-curves
have a vertical tangent at point (A), i.e. dsA = 0. Hence, the fundamental equation
of thermodynamics, see Sect. 12.3.2, in point (A) reads as

T ds = dh − v d p = 0. (21.193)
21.5 Rayleigh Correlation 667

In differential notation the first law of thermodynamics obeys

dh + c dc = δq. (21.194)

A combination with the second law of thermodynamics

δq δψ
ds = + (21.195)
T T


=0

the first law of thermodynamics yields

dh + c dc = T ds = 0. (21.196)

The equation of continuity cρ = const. leads to

d (cρ) = ρ dc + c dρ = 0. (21.197)

A combination of these equation leads to the velocity in point (A), i.e.


 
∂p
c= =a (21.198)
∂ρ s

The velocity at point (A) must therefore be the velocity of sound. Since no restrictions
have been placed on the inlet conditions, it applies to both subsonic and supersonic
inlet states (1).
Following the partial energy equation state (A) can be described with

pA = p1 − ṁ 2
1 (vA − v1 ) . (21.199)

With the conservation of mass it is


cA c1
= = ṁ 1 . (21.200)
vA v1

The velocity in (A) is the velocity of sound, i.e.


 √
cA = κ RTA = κ pA vA . (21.201)

A combination of these equations leads to

p1 + ṁ 2
1 v1
pA = (21.202)
1+κ
668 21 Steady State Flow Processes

respectively
 2 
v1 κ pA
vA = κ pA = 2
→ cA = κ RTA (21.203)
c1 ṁ 1

and
pA vA
TA = (21.204)
R

The caloric state values can then be calculated

TA pA
sA = sref + c p ln − R ln (21.205)
Tref pref

and
h A = h ref + c p (TA − Tref ) (21.206)

Each Rayleigh-curve is a function of an initial state (1) and a mass flux density ṁ 1 ,
i.e. its velocity c1 .

21.6 Normal Shock

Let us now investigate what can happen in a tube flow with a constant cross-section
under supersonic but adiabatic conditions. The considerations in Sect. 21.4 have
shown that the Fanno-correlation must be fulfilled. According to the first law of
thermodynamics, the energy must be constant, i.e.

1 1
h 1 + v12 ṁ 2 = h 2 + v22 ṁ 2 (21.207)
2 2

Since no heat is exchanged, the first law of thermodynamics simplifies to a constant


specific total enthalpy. It has been shown that the flow decelerates due to friction until
Ma = 1 is reached in state (A). As shown in Fig. 21.23, the lower supersonic branch
of the Fanno-curve must therefore be followed. Under this premise, a pressure p2 ,
which is determined e.g. by a huge environment, cannot be reached in case

p2 > pA . (21.208)

However, a so-called normal shock can occur: According to [40] the fluid compresses
in an unsteady change of state, i.e. it is non-quasi-static, from a lower, supersonic
pressure of state (1) to the fixed pressure of subsonic state (2), that is ruled by the
environment for instance. This normal shock (“jump”) is supposed to be linear, i.e.
21.6 Normal Shock 669

Fig. 21.23 Normal shock, adiabatic tube flow

its spatial expansion is neglected. In case the spatial expansion is low, there is no
friction. Thus, for the change of state from (1) to (2) a frictionless tube flow can be
assumed, that needs to follow the Rayleigh correlation. This correlation has shown,
that in case no friction occurs, the total pressure remains constant, i.e.

p2 − p1 + ṁ 2 (v2 − v1 ) = 0 (21.209)

In other words, the states (1) and (2) needs to fulfil both the Fanno- and Rayleigh-
correlation under these particular conditions, cf. Fig. 21.23. Consequently, both
states, i.e. (1) and (2) are on a par. Once, state (2) is reached, the further process
takes place under subsonic conditions and is friction-controlled again.
The change of state from (1) to (2) can be visualised in a h, s-diagram as follows:
• A Fanno-curve is calculated by the given mass flux density and any supersonic
state in front of the normal shock, e.g. state (I).
• State (2) is then unambiguously determined by the Fanno-curve and the isobar p2 .
• A Rayleigh-curve is calculated by the given mass flux density and state (2).
• A second point of intersection between Fanno- and Rayleigh-correlation occurs,
i.e. state (1).
• The normal shock takes place from (1) to (2).
Due to the unsteady jump and its non-quasi-static character, the normal shock
comes along with a rise of entropy, see Fig. 21.23, so it does not violate the second
law of thermodynamics. According to the second law of thermodynamics, entropy
670 21 Steady State Flow Processes

can not be destroyed. This is the reason why a normal shock can not take place from
(2) to (1), as this would require a destruction of entropy, as shown in Fig. 21.23.
The entropy increase has been treated as a discontinuity—otherwise it has not
been possible to get from state (1) to state (2). In reality, however, a normal shock is
distributed over several free path lengths of the particles, so that it can be treated as
a steady effect. The entropy increase can therefore be explained by friction effects
and heat fluxes within the normal shock.

21.7 Supersonic Flows

21.7.1 Flow of a Converging Nozzle

Before discussing supersonic flows in a so-called Laval nozzle, let the focus first
be on a much simpler problem, as sketched in Fig. 21.24. A huge tank contains an
ideal gas in rest, i.e. its velocity is c0 = 0, given by its state values pressure p0 and
temperature T0 . A small nozzle connects the container with the environment (env).
However, the nozzle is converging, i.e. in flow direction its cross-section decreases
with a narrowest cross-section (n) at its end, see Fig. 21.24. In case

p0 > penv (21.210)

there is a flow from the vessel into the environment. Nozzle and vessel are supposed
to be adiabatic. Since the tank is to be very large, its internal state is time-invariant,
so a steady state is to be investigated.

Fig. 21.24 Steady flow from


vessels according to [1]
2.7 Supersonic Flows 671

First Law of Thermodynamics


To derive the outflow velocity cn , the first law of thermodynamics is applied from
state (0) to state (n) with the premises made and yields

1 2 
q0n + wt,0n = 0 = h n + h 0 + cn − c0 . (21.211)
2
Applying the caloric equation of state and taking into account, that c0 = 0 leads to
 
Tn
cn2 = 2c p (T0 − Tn ) = 2c p T0 1 − . (21.212)
T0

In case the change of state is isentropic, i.e. adiabatic and frictionless, it follows for
an ideal gas, that
  κ−1
Tn pn κ
= . (21.213)
T0 p0

Thus, the outflow velocity is


   κ−1 
pn κ
cn2 = 2c p (T0 − Tn ) = 2c p T0 1 − . (21.214)
p0

The tank temperature T0 follows from the thermal equation of state, i.e.
p0 v0
T0 = (21.215)
R
and it further is κ
cp = R. (21.216)
κ −1

This finally leads to the outflow velocity


  
   κ−1
 κ pn κ

cn = 2 p0 v0 1 − (21.217)
κ −1 p0

In case the flow would not be frictionless, i.e. non-isentropic, the outflow velocity
can be derived with a polytropic change of state with n < κ and leads to
  
   n−1
 κ p
= 2
n
n
cn,fric p0 v0 1 − < cn . (21.218)
κ −1 p0
672 21 Steady State Flow Processes

However, the focus is now on the ideal case of an isentropic change of state as a
benchmarking change of state.
Conservation of Mass
The continuity equation for any cross-section of the nozzle, i.e. also in the narrowest
cross-section, in steady state reads as

cA cn An
ṁ = = = ṁ n = const. (21.219)
v vn

Substituting the velocity cn of the narrowest cross-section brings


  
   κ−1
An 
2 κ p n
κ
ṁ n = p0 v0 1 − . (21.220)
vn κ −1 p0

Due to the isentropic change of state it is


  κ1
p0
vn = v0 . (21.221)
pn

Hence, it follows accordingly


  
  κ1 
 κ p0   κ−1
pn 2 pn κ
ṁ n = An 1− . (21.222)
p0 κ − 1 v0 p0

Now, a so-called flow function


n for the narrowest cross-section is defined, so that

p0 
ṁ n = An
n 2 = An
n 2ρ0 T0 . (21.223)
v0

Thus, the dimensionless flow function is


   1
   κ−1
 κ pn κ pn κ

n =  1− (21.224)
κ −1 p0 p0

respectively with the pressure ratio


pn
πn = (21.225)
p0

it is 
κ  κ−1  1

n = 1 − πn κ πnκ . (21.226)
κ −1
2.7 Supersonic Flows 673

Since the mass flow rate needs to be constant in steady state, a flow function can be
defined at any position within the nozzle, i.e. from state (0) to (n), so that the general
notation is

p0 
ṁ n = ṁ = A
2 = A
2ρ0 T0 (21.227)
v0

with

κ  κ−1
 1

=
(π ) = 1−π κ πκ (21.228)
κ −1

and
p
π= (21.229)
p0

Now, the maximum mass flow rate that can occur in a converging nozzle is investi-
gated. Unfortunately, the mathematical function of the converging shape, i.e. the local
cross-section A, is not given. Anyhow, the local flow function
, see Eq. 21.228, can
be plotted over the local pressure ratio π , cf. Fig. 21.25. Since the mass flux needs
to be constant in steady state, the corresponding cross-section can easily be added
qualitatively in the plot as well, following

ṁ 1
A= √ ∝ . (21.230)

2ρ0 T0

Thus, the larger the local


, the smaller the local cross-section A. The flow starts
within the tank in state (0), i.e. at π = 1 at the right position of Fig. 21.25. On

Fig. 21.25 Local flow function


for κ = 1.4, state (0) in rest
674 21 Steady State Flow Processes

the way to (n), the cross-section becomes smaller because the nozzle is assumed to
converge. Correspondingly, the flow function
rises and reaches a maximum. Due
to the constant mass flow rate, the cross-section then needs to have a minimum, that
has been introduced as An . A further movement to the left beyond that point is not
possible, since then the nozzle would need to diverge, i.e. the dashed lines can not
be reached with a converging geometry. The pressure ratio at (n) can be calculated
by finding the maximum of
(π ), i.e.
1
d
2π κ − (κ + 1) π
=  κ−1 = 0. (21.231)
dπ κ
2 (κ − 1) π 1−κ π κ −1

Thus, the maximum, which is at the narrowest cross-section, of the curve is at


  1−κ
κ
κ +1
πmin = . (21.232)
2

The maximum flow function at that point follows according to


      1
κ 2 κ + 1 1−κ

max = 1− (21.233)
κ −1 κ +1 2

respectively
 1 
 κ−1
2 κ

max = . (21.234)
κ +1 κ +1

However, that point is called critical point of the nozzle18 and its state values then
follow   κ   κ
κ + 1 1−κ κ + 1 1−κ
pcrit = p0 → πcrit = (21.235)
2 2

and due to the isentropic change of state


 − κ1   κ−1
1
pn κ +1
vcrit = v0 = v0 (21.236)
p0 2

respectively
  κ−1
κ
pn 2
Tcrit = T0 = T0 . (21.237)
p0 κ +1

18For a given p0 it is not possible to reach a pressure smaller than pcrit in a converging nozzle, see
Fig. 21.25.
2.7 Supersonic Flows 675

Since the state values for the critical state are now known, the velocity can be calcu-
lated according to Eq. 21.217
  
   κ−1
 κ p κ
= 2
crit
ccrit p0 v0 1 − . (21.238)
κ −1 p0

With the other critical state values derived previously, one obtains
√ 
ccrit = κ pcrit ccrit = κ RTcrit = acrit . (21.239)

Theorem 21.7 The maximum velocity at the narrowest cross-section of a converging


nozzle is therefore the velocity of sound, i.e. Ma = 1, when the pressure reaches its
critical value pcrit . The pressure cannot fall below this critical value.
In Fig. 21.25 it has been assumed that π can freely vary from 0 . . . 1 and the
consequences have been investigated mathematically. From a thermodynamic point
of view, the range of π has been further limited by πcrit , in the best case at the
narrowest cross-section, to 1 in state (0). The range from 0 . . . πcrit is impossible
for a converging contour. Now several case studies are discussed, focusing on what
happens when πcrit is not reached in the narrowest cross-section, e.g. due to ambient
counterpressure. Figures 21.26 and 21.27 therefore show cases (a), (b), (c) and (d),
which need to be discussed further and which show that not the whole range of π can
be covered. It is known that the cause of any flow is a pressure difference: The cause

Fig. 21.26 Converging nozzle scenarios I/II


676 21 Steady State Flow Processes

Fig. 21.27 Converging nozzle scenarios II/II

of flow through the converging nozzle is the pressure potential between vessel p0
and ambient pressure penv , which acts as a kind of counterpressure. Let us therefore
gradually increase the pressure p0 in the vessel and fix the ambient pressure:

(a) The pressure in the vessel p0 is greater than ambient pressure. As shown in
Fig. 21.27 the flow starts, but the critical pressure at the narrowest cross-section,
i.e. at the outlet is not reached, i.e.
   κ
pn penv  pcrit κ + 1 1−κ
= > = . (21.240)
p0 p0 (a) p0 2

The pressure at the outlet (n) is ambient pressure. The according outlet veloc-
ity, see Mach-number in Fig. 21.27, is smaller than the velocity of sound.
Figure 21.26 additionally shows the corresponding contour of the cross-section
as a function of π . At the narrowest cross-section An the flow function reaches


n,(a) <
max . (21.241)

Thus, the maximum mass flow rate corresponding to the contour of the converg-
ing nozzle and the vessel pressure p0 has not yet been reached.
(b) As the pressure in the tank is increased further, the velocity increases due to
the greater pressure potential which drives the flow. In case (b) the chosen tank
pressure causes the outlet pressure at the narrowest cross-section to be
   κ
pn penv  pcrit κ + 1 1−κ
= = = . (21.242)
p0 p0 (a) p0 2
2.7 Supersonic Flows 677

Consequently, the velocity reaches its maximum value, i.e. Ma = 1, see


Fig. 21.27. According to Fig. 21.26, the flow function
is maximised at the
narrowest cross-section, i.e.

n,(b) =
max . (21.243)

Therefore, the mass flow rate is maximum and cannot be exceeded for the given
condition (0) and geometry of the nozzle. The outlet pressure is equal with
ambient pressure. This case is the design case of a perfect converging nozzle.
(c) In case the tank pressure p0 is further increased, the critical pressure pcrit rises
as well, but according to Fig. 21.25 the critical pressure ratio is still fixed to
  1−κ
κ
pcrit κ +1
πcrit = = . (21.244)
p0 2
 
Thus, the dimensionless function π = f Lx is the same as in case (b). Equal to
case (b), the flux function reaches its maximum value, i.e.


n,(c) =
max . (21.245)

Consequently, the Mach-number at the outlet still is Ma = 1. However, ambient


pressure is not reached at the outlet, i.e.

pn = pcrit > penv . (21.246)

According to [1], the nozzle is still exited with

pn = pcrit . (21.247)

As it exits the nozzle, the flow begins to expand periodically and eventually
reaches ambient pressure. However, let us take a look at the mass flow rate:
According to Eq. 21.223, it follows

p0
ṁ n = An
n 2 . (21.248)
v0

The narrowest cross-section and the local flow function at this point are the
same as in (b). However, the pressure p0 has increased so that v0 decreases if the
temperature T0 of the tank has not changed compared to (b), i.e.

RT0
v0 = . (21.249)
p0

Consequently, the mass flow rate is greater than in (b), but still maximum for the
given geometry and initial pressure p0 .
678 21 Steady State Flow Processes

(d) However, a smaller πn < πcrit is not possible, as this would require a further
increase in the cross-section. Yet this is in contradiction to a converging nozzle.
This case is shown in Fig. 21.26.
The mass flow rate through the nozzle in critical state is maximum, see case (b)
respectively (c) in Figs. 21.26 and 21.27. The mass flow rate is proportional to the
narrowest cross-section An and its flow function at that point
n , see Eq. 21.230.
The cross-section at the narrowest point is fixed, but only in case (b) respectively (c),
the flow function reaches its maximum.

The curves in Fig. 21.27 have been calculated by solving the following equations:
1. The first law of thermodynamics for adiabatic flows, i.e.

1 1 1
h + = const. = h + c2 = h 0 = h 1 + c12 = h n + cn2 . (21.250)
2 2 2
2. Isentropic change of state, i.e.

pvκ = p0 v0κ = p1 v1κ = pn vnκ . (21.251)

3. Conservation of mass, i.e.

cA c1 A1 cn An
= = . (21.252)
v v1 vn

4. Thermal equation of state, i.e.


pv = RT. (21.253)

5. Caloric equation of state, i.e.

h = c p (T − Tref ) + h ref (21.254)

6. Second law of thermodynamics,19 i.e.

dT dp
ds = 0 = c p ln − R ln . (21.255)
T p

21.7.2 Laval-Nozzle

In the previous section it has been shown that the maximum outlet velocity from a ves-
sel through a converging nozzle is the velocity of sound when the pressure in the nar-
rowest cross-section reaches its critical value pcrit . This critical value is fluid specific

19 The flow is adiabatic and reversible, i.e. it is isentropic.


2.7 Supersonic Flows 679

and has been derived in the previous section—depending on the resting pressure p0
in the tank and the counterpressure, this critical value can be reached. In this section,
it is now examined how a supersonic state, i.e. Mach numbers greater than one, can
be achieved. For this purpose, a comprehensive theoretical approach is followed, i.e.
a correlation between cross-sectional area and mass flux density through a nozzle is
investigated in a differential approach. Due to the conservation of mass in a steady
state nozzle with variable cross-section, the mass flow rate must be constant, i.e.

ṁ = cρ A = const. (21.256)

The derivation yields


d (cρ A) = 0. (21.257)

Rearranging leads to
A d (cρ) + ρc dA = 0. (21.258)

Hence, the change of cross-section follows

dA d (cρ) c dρ + ρ dc dρ c dc
=− =− =− − 2 . (21.259)
A ρc ρc ρ c

Equation 21.259 can also be written with the mass flux density, i.e.


ṁ  = = ρc (21.260)
A
so that
dA dṁ 
= −  . (21.261)
A ṁ

A physical interpretation20 is, that the cross-sectional area A has a minimum, i.e. the
first derivative is dA = 0 and the second derivative is d2 A > 0, at the point where the
mass flow density ṁ  is maximum, i.e. first derivative is dṁ  = 0 and second deriva-
tive is d2 ṁ  < 0. This narrowest cross-section is important in the further course.
However, the first derivative at this location is

dA dṁ 
= −  = 0 (21.262)
A ṁ
and the second derivative at this location is

d2 A d2 ṁ 
=− 
. (21.263)
A

 ṁ


>0 <0

20 Simplified, it can be said that at constant mass flux, the mass flux density must become greatest
at the narrowest cross-section.
680 21 Steady State Flow Processes

This interpretation of a maximum mass flux density at the narrowest cross-section


is congruent with the flow function
introduced in Sect. 21.7.1, cf. e.g. Fig. 21.25.

Assuming an adiabatic and frictionless, i.e. isentropic, flow with negligible changes
of potential energies, the partial energy equation for technical work21 reads as

2
1 2 
wt = 0 = v dp + c2 − c12 . (21.264)
2
1

In differential notation it yields

v d p + c dc = 0. (21.265)

A combination with the equation of continuity, see Eq. 21.259, results in

dA dρ v dp
=− + 2 . (21.266)
A ρ c

In the derivation of the velocity of sound, it has already been demonstrated for an
isentropic change of state, see Eq. 21.99, that

dp
dρ = . (21.267)
a2
Hence, the following correlation can be found for the cross-sectional area A of the
nozzle  
dA 1 1
= − v dp (21.268)
A c2 a2

Introducing the Mach-number


c
Ma = (21.269)
a
and applying Eq. 21.265 lead to the so-called Rankine–Hugoniot equation, i.e.

dc dA
=−   (21.270)
c A 1 − Ma2

Respectively, with Eq. 21.265 it is

dp dA
=   (21.271)
ρc 2 A 1 − Ma2

21 This is equivalent with a constant total pressure. Note that no dissipation occurs.
2.7 Supersonic Flows 681

Fig. 21.28 Overview subsonic/supersonic flows

The consequences of these two equations are shown in Fig. 21.28:


• Subsonic flows, i.e. Ma < 1

According to the Rankine–Hugoniot equation, a nozzle, i.e. the acceleration of a


fluid such that dc > 0, requires a converging geometry, i.e. dA < 0. The pressure
while passing the nozzle decreases. A diffuser, in contrast, which decelerates a
fluid so that dc < 0, requires a diverging geometry with dA > 0. The pressure
within the diffusor rises.

• Supersonic flows, i.e. Ma > 1

According to the Rankine–Hugoniot equation, a nozzle, i.e. dc > 0, requires a


divergent geometry, i.e. dA > 0. The pressure while passing the nozzle decreases.
A diffuser characterised by dc < 0, in contrast, requires a converging geometry
with dA < 0. The pressure within the diffusor rises.
This knowledge can be applied to derive the generation of a supersonic flow with
subsonic inlet conditions: A converging/diverging nozzle shape is needed, which is
shown in Fig. 21.29. This nozzle type is a so-called Laval nozzle. The set-up shall
be the same as before: a container contains an ideal gas at rest (0), i.e. the pressure
shall be p0 , the temperature T0 and the velocity c0 = 0. The other state values follow
accordingly, while the potential energy being ignored. State (1) is the inlet state in
the converging part of the Laval-nozzle, state (n) is the state at the narrowest cross-
section and state (2) indicates the outlet state. Outside shall be ambient pressure
682 21 Steady State Flow Processes

Fig. 21.29 Steady flow from vessels using a Laval-nozzle

Fig. 21.30 Pressure characteristic Laval-nozzle

penv . Obviously, the driver for a flow is the pressure potential between tank and
environment. In case ambient pressure is equal to the pressure within the tank, no
flow develops. Let us now start to increase the pressure p0 in the tank while fixing
ambient pressure penv . The scenarios (a)–(h) are shown in Figs. 21.30 and 21.31.
2.7 Supersonic Flows 683

Fig. 21.31 a Venturi-nozzle, f Supersonic flow

It should be noted that it is not sufficient to consider only the pressure difference
between the tank and the environment, but that the level of ambient pressure must
also be taken into account, as this controls the flow in the Laval nozzle:
(a) The pressure within the tank p0 is greater than the ambient pressure penv . A flow
develops, but the pressure in state (n) is smaller than the critical pressure, i.e.

πn < πcrit . (21.272)

The flow leaves with


π2 = πenv → p2 = penv . (21.273)

The counterpressure penv is too large to reach a supersonic state within the nozzle.
Velocity of sound is not reached and a pressure distribution as shown in Fig. 21.30
is measured. The flow leaves with subsonic velocity and the mass flow rate is
not maximised according to the geometry of the Laval-nozzle. The temperature
profile as well as the Mach-number profile are additionally shown in Fig. 21.31.
(b) The same as in (a) while pn further approximates pcrit .
(c) The same as in (b) while pn further approximates pcrit .
(d) The same as in (c) while pn further approximates pcrit .
(e) In this case, the mass flux is sufficiently high, so that

πn = πcrit . (21.274)
684 21 Steady State Flow Processes

However, it is

πn = πcrit < π2 = πenv → pn = pcrit < p2 = penv . (21.275)

The counterpressure ratio πenv has been lowered compared to the previous cases
but it is still too high for a supersonic flow in the diverging sector. According to
[1] a normal shock, cf. Sect. 21.6, takes place at the narrowest cross-section.
(f) If the design fits to the geometry (“adapted Laval-nozzle”) and the initial condi-
tion in the tank, i.e. state (0), the outlet pressure is ambient pressure, i.e.

π2 = πenv → p2 = penv (21.276)

and the pressure in the narrowest cross-section fulfils

πn = πcrit > π2 = πenv → pn = pcrit > p2 = penv . (21.277)

The mass flow rate is maximised for p0 and the geometry of the nozzle. Note
that in this case π2 is purely a function of the geometry of the nozzle.
(g) A supersonic state is reached within the diverging part of the nozzle, i.e.

πn = πcrit → pn = pcrit . (21.278)

However, the ambient pressure information can not get upstream, since the flow
moves faster than velocity of sound. In this case, the ambient pressure is in the
range

πenv,(g) = πenv,(e) . . . πenv,(f) → penv,(g) = penv,(e) . . . penv,(f) . (21.279)

Normal shocks occur in the nozzle, i.e. the fluid is compressed unsteadily from
the lower pressure to the higher pressure in the environment. Note that this
characteristic within the nozzle is not sketched in Fig. 21.30.
(h) A supersonic state is reached within the diverging part of the nozzle. Same as
before, the ambient pressure information can not get upstream, since the flow
moves faster than velocity of sound. However, the pressure at the outlet p2 can
be greater than the ambient pressure penv , i.e.

π2 > πenv → p2 > penv , (21.280)

while the critical pressure at (n) is still reach. It further is

πenv < πenv,f) . (21.281)

The fluid starts to expand as soon as it leaves the nozzle. Periodic jet expansions
and constrictions occur, see [1]. However, this scenario may occur if p0 is further
increased beyond the design case according to (f). Any further increase of p0
2.7 Supersonic Flows 685

leads to the same dimensionless characteristic


p x
π= = f (21.282)
p0 L

as shown in Fig. 21.30f. This function depends exclusively on the geometry of


the Laval nozzle, see Fig. 21.25, i.e. the same πcrit and the same π2 are reached.
Increasing the pressure within the tank  p0 leads to a greater mass flow rate.
Although the function π = pp0 = f Lx is fixed for supersonic flows, the ratio
penv
p0
decreases with increasing p0 at constant ambient pressure.
A more detailed explanation regarding the scenarios can be found in [1, 38]. The
shown graphs have been calculated with the same approach as in Sect. 21.7.1, i.e.
solving Eqs. 21.250–21.254. Figure 21.31 compares cases (a), i.e. a subsonic flow,
with the design case (f) of a Laval-nozzle. It shows the dimensionless shape of the
Laval-nozzle, as well as the dimensionless temperature and pressure distributions.
In case (f) the critical state at the narrowest cross-section is reached, so that
  1−κ
κ
κ +1
pcrit = p0 (21.283)
2

respectively
  κ−1
κ
pn 2
Tcrit = T0 = T0 . (21.284)
p0 κ +1

Consequently, cf. Eq. 21.239, the Mach-number is

Ma = 1. (21.285)

For the example shown with the geometry applied, the pressure ratio at the outlet
meets p2
π2 = = 0.0712. (21.286)
p0

Case (a) is also called a Venturi-nozzle. The entire flow remains subsonic but is
supposed to be adiabatic and frictionless, i.e. isentropic. It should therefore not be
confused with an adiabatic throttle, which has a similar geometric shape but relies
on friction to reduce pressure.
Adiabatic/Polytropic Laval-Nozzle
Figure 21.32 shows a comparison between an isentropic Laval nozzle, as studied
so far, and an adiabatic/polytropic, i.e. frictional, Laval nozzle. In this case, the
algorithm for calculating this flow is the same as before, see Eqs. 21.250–21.254,
but the change of state is treated as polytropic with n < κ, see Eq. 21.251, i.e.

pvn = p0 v0n = p1 v1n = pn vnn = p2 v2n . (21.287)


686 21 Steady State Flow Processes

Fig. 21.32 Supersonic exit Laval-nozzle a isentropic, b polytropic/adiabatic, i.e. with friction

Furthermore it is
ds > 0. (21.288)

As can be seen from the diagram in Fig. 21.32, the supersonic flow does not reach
its critical pressure in the narrowest cross-section, but in the diverging part of the
nozzle. The Ma-number follows accordingly and thus is shifted to the right, i.e.
Ma = 1 is in the diverging section. Figure 21.33 shows the isentropic as well as
the frictional Laval-nozzle in a T, s-diagram. Obviously, the polytropic change of
state with n < κ causes the specific entropy to rise. This increase correlates with
the friction, that can be visualised by the area beneath the change of state in the
T, s-diagram. The temperature in the frictional case is greater than for isentropic
conditions.
Laval-Diffusor
If the convergent/divergent shape of a Laval nozzle is operated with a supersonic
inlet, cf. Fig. 21.34, the characteristic follows a diffusor. According to the Rankine–
Hugoniot-equation
dc dA
=−   (21.289)
c A 1 − Ma2

the supersonic flow is decelerated in the converging part of the geometry. Under
isentropic conditions the flow reaches Ma = 1 in the narrowest cross-section and
2.7 Supersonic Flows 687

Fig. 21.33 Supersonic exit Laval-nozzle, T, s-diagram

Fig. 21.34 Laval-diffusor


688 21 Steady State Flow Processes

then further decelerates in the diverging part. Finally, the flow leaves with a greater
pressure as at the inlet, while the velocity is smaller. Note that dimensionless pressure
and temperature in Fig. 21.34 are referred to the inlet state of the nozzle (1). A state
(0), which used to be the resting state of a tank, no longer makes sense, since state
(1) is now given and the system has an initial velocity.
Chapter 22
Thermodynamic Cycles with Phase
Change

Although thermodynamic cycles have already been introduced in Sect. 17.2, this
chapter deals with cycles in which the working fluid is subject to phase changes.
When a system finally reaches the initial state after several changes of state and the
process starts all over again, this is called a thermodynamic cycle. Thus, all state
values reach their initial value after passing the cyclic process, i.e.

dZ = 0. (22.1)

A distinction has been made regarding the application of such processes: Clockwise
cycles convert heat into work, and with regard to the second law of thermodynamics,
a thermal efficiency of 100% is not possible, i.e. heat needs to be passed to the
environment. A typical representative of such a cycle is the Joule process. However,
the working fluids so far used to be ideal gases. In this chapter the focus is on
clockwise cycles with working fluids, that typically change their aggregate state, e.g.
a steam power process. Counterclockwise cycle shift heat from a lower temperature
to a higher level. In order to do so, according to the second law of thermodynamics,
work needs to be supplied. This operating principle can be applied to both heat pumps
and cooling machines. However, these cycles have only been briefly discussed in Part
I, so they are the focus of this chapter.
The Carnot-process, cf. Sect. 15.3, is an idealised benchmarking cyclic process,
that consists of two reversible isothermal and two reversible adiabatic, i.e. isentropic,
changes of state. Thus, every change of state is reversible and the process reaches
its maximum efficiency. This process has already been discussed in Part I and is
exemplified in Fig. 22.1 for a clockwise cycle. The process consists of the following
changes of state:
• (1) → (2): Isothermal expansion
• (2) → (3): Reversible, adiabatic expansion

© Springer Nature Switzerland AG 2022 689


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_22
690 22 Thermodynamic Cycles with Phase Change

Fig. 22.1 Clockwise Carnot process: thermal engine

• (3) → (4): Isothermal compression


• (4) → (1): Reversible, adiabatic compression
The thermal efficiency of such a process depends exclusively on the minimum and
maximum operating temperature, i.e.

Tmin
ηC = 1 − (22.2)
Tmax

22.1 Steam Power Process

22.1.1 Clausius–Rankine Process

The Clausius–Rankine process is an ideal1 reference cycle for steam power plants,
i.e. it covers the phase changes of the working fluid that usually take place in the
liquid and gaseous state. Its principle layout is shown in Fig. 22.2 and is similar to
the Joule process already presented, see Part I. The process consists of the following
changes of state, that are illustrated in a p, v-as well as T, s-diagram, see Fig. 22.3:
• (1) → (2): Adiabatic, frictionless compression

The required compression to reach a high pressure level is idealised and supposed to
be adiabatic and frictionless. It is advantageous to increase the pressure of a liquid
fluid instead of a gaseous fluid, since the technical work under these conditions,
neglecting the kinetic and potential energies, is as follows

2
wt,12 = v d p > 0. (22.3)
1

1 Idealised in this case means, that there is no generation of entropy.


22.1 Steam Power Process 691

Fig. 22.2 Clausius–Rankine cycle: sketch

Fig. 22.3 Clausius–Rankine cycle: a p, v-diagram, b T, s-diagram

In contrast to gases/steams the specific volume v of liquids is rather small. The


first law of thermodynamics obeys

wt,12 = wt,P = h 2 − h 1 > 0. (22.4)

In the example shown, it is assumed that state (1) is a ( ) state.


• (2) → (3): Isobaric heat transfer + phase change
If the supply of heat is isobaric, the partial energy equation with the same premise
as for the change of state from (1) to (2) is as follows

3
wt,23 = v d p +ψ23 = 0. (22.5)
2
  
=0
692 22 Thermodynamic Cycles with Phase Change

Thus, no dissipation takes place, so that this change of state is also idealised. The
increase in specific entropy is only due to the supplied heat. The initially liquid
fluid becomes a saturated liquid in the ( ) state, vaporises completely and is further
superheated to reach the (3) state. The first law of thermodynamics yields

q23 = q = h 3 − h 2 > 0. (22.6)

Due to the chosen system boundary,2 the entropy generation associated with the
heat transfer from the environment to the fluid is not part of the system, cf.
Sect. 17.1.5 for details.
• (3) → (4): Adiabatic, frictionless expansion
In this step, the fluid, which is at a high pressure level and, due to heating, at a
high energetic level, now releases technical work in the adiabatic and frictionless
turbine, so that the pressure level decreases. The first law of thermodynamics,
which ignores both kinetic and potential energies, reads as follows
wt,34 = −wt,T = h 4 − h 3 < 0. (22.7)

The partial energy equation follows


4
wt,34 = v d p < 0. (22.8)
3

• (4) → (1): Isobaric heat transfer + phase change


If the heat release is isobaric, the partial energy equation under the same conditions
as for the change of state from (1) to (2) is as follows
1
wt,41 = v d p +ψ41 = 0. (22.9)
4
  
=0

Thus, no dissipation occurs, so that this change of state is idealised as well. The
wet steam condenses and finally reaches state (1). The first law of thermodynamics
follows
q41 = −q0 = h 1 − h 4 . (22.10)

State (4) is assumed to have reached the wet steam region after the turbine.
The graphs in Fig. 22.3 prove, that the Clausius–Rankine cycle is a clockwise cycle.

2 Only the flux of the fluid through the heat exchanger is part of the system boundary. The envi-
ronment, which is at a higher temperature level and thus forms a discontinuity with the fluid, is not
included.
22.1 Steam Power Process 693

Work is therefore effectively released, i.e. the work released by the turbine is
greater than the work supplied at the water pump, i.e.

|wt,34 | > |wt,12 |. (22.11)

The overall first law of thermodynamics obeys


 
wt + q = 0 → wt,12 + wt,34 + q23 + q41 = 0. (22.12)
     
=weff =qeff

The effective work


weff = wt,12 + wt,34 (22.13)

is visualised as enclosed area in the p, v-diagram.3 Since the entire process takes
place without entropy generation, it is the specific heat that is represented as the area
under each change of state in the T, s-diagram. Consequently, the enclosed area in
the T, s-diagram represents the effective heat

qeff = q23 + q41 . (22.14)

According to the first law of thermodynamics, this is equivalent to

weff = −qeff . (22.15)

22.1.2 Steam Power Plant

In this section, a non-ideal process is discussed which, in contrast to the Clausius–


Rankine process, generates entropy. Beforehand, the supply of heat in a steam power
process is formally divided, cf. Fig. 22.4. In the first step, i.e. from state (2), preheat-
ing takes place until the subcooled fluid reaches state (3) = ( ). In the second step,
the fluid vaporises completely from state (3) to state (4) = ( ). Finally, superheat-
ing occurs from state (4) to state (5). The corresponding T, s-diagram is shown in
Fig. 22.5. As the specific entropy in the pump and turbine increases, the process is
not ideal, i.e. it differs from a perfect Clausius–Rankine process. It is advantageous
to represent a steam power process in an h, s-diagram, as has been done in Fig. 22.6,
since the specific work as well as the specific heat can be easily determined under the
premise that both potential and kinetic energies are neglected. It is therefore further
assumed that the pump and turbine are adiabatic. For the turbine, the first law of
thermodynamics is exemplified as follows

wt,56 + q56 = h 6 − h 5 < 0. (22.16)



=0

3 Since potential and kinetic energies have been ignored. Furthermore, no dissipation occurs.
694 22 Thermodynamic Cycles with Phase Change

Fig. 22.4 Steam power process: boiler

Fig. 22.5 Steam power process: T, s-diagram

For the overall heat supply it yields

wt,25 +q25 = h 5 − h 2 > 0. (22.17)



=0
22.1 Steam Power Process 695

Fig. 22.6 Steam power process: h, s-diagram

Assuming that the turbine is adiabatic, the increase in specific entropy from (5)
to (6) is due to the frictional entropy generation, see Fig. 22.6, i.e.

ds = δsi + δsa → s56 = s6 − s5 = si,T . (22.18)



=0

In order to calculated a steam power process, the required caloric state values can
be taken from steam tables. However, the expansion shown from state (5) to state
(6) is not perfectly chosen, as the expanded vapour is still at a high temperature
respectively enthalpy level. Therefore, it is very unlikely that the fluid expands into a
superheated state, but into the wet steam region. According to the h, s-diagram, see
Fig. 22.6, the lower pressure level p1 is further decreased. Typical steam parameters
are ϑ5 = 550 ◦ C and p5 = 250 bar. The work effectively gained in the plant is

wt,eff = wt = wt,12 + wt,56 . (22.19)

Technical work needs to be supplied to the pump, whereas the turbine releases tech-
nical work. The first law of thermodynamics for the overall power plant leads to
 
wt + q = 0. (22.20)

Thus, according to Fig. 22.4 it is

wt,eff = −(qin − |qout |). (22.21)


696 22 Thermodynamic Cycles with Phase Change

With the introduction of the isentropic efficiency, see Sect. 17.1, the first law of
thermodynamics for the adiabatic and irreversible pump results as follows

h s
wt,12 = h 2 − h 1 = (22.22)
ηs,P

and for the adiabatic, irreversible turbine

wt,56 = h 6 − h 5 = h s ηs,T . (22.23)

Applying the first law of thermodynamics for the isobaric boiler brings

qin = q25 = h 5 − h 2 . (22.24)

The condenser is also treated isobarically, so that the first law of thermodynamics is

qout = q61 = h 1 − h 6 . (22.25)

Hence, it is now possible to calculate the thermal efficiency of the steam power
plant, i.e.
Benefit |wt,eff |
ηth = = . (22.26)
Effort qin

Substitution leads to
|h 2 − h 1 + h 6 − h 5 |
ηth = . (22.27)
h5 − h2

The technical work wt,12 supplied to the pump is often neglected compared to the
heat released/supplied as well as to the technical work in the turbine, so that h 2 ≈ h 1 ,
see Fig. 22.6. Thus, the thermal efficiency under these conditions obeys

h5 − h6
ηth = (22.28)
h5 − h1

Reheating

Figure 22.5 has shown that it is advantageous not to let the turbine expand into the
superheated region, but to bring the fluid into the wet steam state. In this way, the
technical work released can be maximised, see Fig. 22.7. However, if the vapour
ratio x is small, the wet steam carries liquid droplets that can cause erosion when
they hit the turbine blades. According to Fig. 22.7, this threat can be avoided by
further shifting state (5) to higher temperatures. However, as mentioned before, the
maximum temperature is limited and an increase in the lower operating pressure,
i.e. bringing state (6) closer to the x = 1 curve, would reduce the performance of
22.1 Steam Power Process 697

Fig. 22.7 Steam power process—expansion in the wet-steam region

Fig. 22.8 Steam power process—reheating

the turbine. A solution is the so-called reheating, as shown in Fig. 22.8: After partial
expansion in a first high-pressure turbine, the fluid is reheated in a heat exchanger.
To achieve the lower operating pressure, the fluid then expands in the low-pressure
turbine. The entire cycle follows:
698 22 Thermodynamic Cycles with Phase Change

Fig. 22.9 Steam power process—reheating

• (1) → (2): Polytropic compression


As in a regular steam power process, the pressure of the liquid fluid is increased. In
most cases, it is assumed that the pump is adiabatic, so that the increase in entropy
is caused by frictional effects. The change of state is therefore polytropic.
• (2) → (3): Preheating
By supply of heat, the supercooled4 liquid is heated to reach the ( ) state.
• (3) → (4): Vaporisation
Due to further supply of heat, the fluid changes from the ( ) to the ( ) state, i.e.
from the left x = 0 curve to the right x = 1 curve of the wet steam region.
• (4) → (5): Overheating
As soon as the vaporisation is complete, the further supply of thermal energy leads
to a sensible heating, i.e. the steam temperature rises.
• (5) → (6): High pressure expansion
The fluid expands from the maximum operation pressure pHD to a medium pressure
pMD . Thus, the minimum operation pressure of the process is not yet achieved in
the HP-turbine. State (6) is superheated fluid.
• (6) → (7): Isobaric reheating
The superheated fluid (6) is further heated by external supply of heat. Figure 22.9
shows, that this change of state is isobaric, i.e. if kinetic and potential energies are

4 Supercooled means, that the fluid temperature is lower than the boiling temperature at this pressure.
22.1 Steam Power Process 699

ignored, it is frictionless and thus idealised. In conventional steam power plants,


however, a pressure drop occurs. As desired, state (7) is now further away from
the x = 1 curve.
• (7) → (8): Low pressure expansion
In this step, the last expansion takes place in the LP turbine to reach the minimum
operating pressure. As state (7) has been moved further to the right, the expansion
can now be carried out without the risk of drop erosion occurring on the turbine
blades.
• (8) → (1): Condensation
In the last change of state, the fluid is now condensed to reach state (1) again.
The thermodynamic cycle is closed. Again, it is assumed that the condensation is
isobaric. The same has been done for the supply of heat from (2) to (5).
Figure 22.9 shows the advantage of the reheating: Erosion-related damage to the
turbine blades can be prevented. Furthermore, the thermal efficiency of such a steam
process is increased as the entire heat qin,1 + qin,2 is supplied at a higher average
temperature level. This phenomenon is called carnotisation. According to the ideal
Carnot process, thermal efficiency is simply a function of the minimum and maximum
operating temperature: the higher the temperature at which heat is supplied, the
greater the thermal efficiency. This is exactly what is utilised in a steam process
with reheating, as the specific heat qin,2 is supplied at a high temperature level. The
thermal efficiency of the process shown, neglecting the rather low technical work of
the pump, is
Benefit |wt,56 + wt,78 |
ηth = = . (22.29)
Effort q25 + q67

Regenerative Feed Water Preheating

The idea of the carnotisation can be further developed. Figure 22.10a shows a steam
power process as discussed before in a T, s-diagram. Obviously, heat is supplied
from state (2) to state (5). Consequently, the averaged temperature T Q̇ in , at which
heat is supplied externally, needs to be somewhere in-between temperature T2 and
temperature T5 . According to Carnot this temperature should be as high as possi-
ble. Unfortunately, it is the preheating that reduces the averaged temperature of the
external supply of heat, as compared to states (3), (4) and (5), the temperature T2 is
rather small. To maximise thermal efficiency, only the heat supplied from outside is
relevant, so the idea of a carnotisation is to realise feed water preheating within the
process, i.e. without transferring thermal energy from outside into the process. The
required energy is therefore taken from the process itself, i.e. regeneratively. The
major advantage is shown in Fig. 22.10b: The averaged temperature T Q̇ in , at which
external heat is supplied, rises and increases the thermal efficiency.
700 22 Thermodynamic Cycles with Phase Change

Fig. 22.10 Steam power process—a No feed water preheating b Regenerative feed water preheat-
ing, cf. Fig. 22.11

Fig. 22.11 Steam power process—regenerative feed water preheating

Figure 22.11 shows, how regenerative5 feed water preheating from state (1) to state
(3) can be realised. Partially expanded vapour is extracted from the turbine, state (6),
which is still at a high temperature level. This extracted mass flux can therefore be
used to heat the feed water after the first water pump, state (1). This first water pump
must raise the pressure of state (8) to the same pressure as the flow taken from the

5 The required energy comes from the process itself.


22.1 Steam Power Process 701

turbine (6). The first law of thermodynamics for the adiabatic preheater yields6 :

ṁ 1 h 1 + ṁ 6 h 6 = (ṁ 1 + ṁ 6 ) h 2 . (22.30)
  
=ṁ

The second law of thermodynamics reads as:

ṁ 1 s1 + ṁ 6 s6 + Ṡa + Ṡi = (ṁ 1 + ṁ 6 ) s2 . (22.31)


   
=0 =ṁ

Thus, the entropy generation due to the mixing is

Ṡi = ṁ 1 (s2 − s1 ) + ṁ 6 (s2 − s6 ) . (22.32)

After leaving the preheater, cf. state (2)7 in Fig. 22.11, the pressure in a second
water pump must be further increased to state (3). The sketched system boundary
in Fig. 22.11 clearly shows that water preheating is now achieved by internal means
and no external energy is transferred across the system boundary to heat the water.
However, the externally supplied heat flux Q̇ in is required for both vaporisation and
superheating of the fluid.

22.2 Heat Pump and Cooling Machine

In this section, the focus is on counterclockwise cycles, i.e. thermodynamic cycles


in which heat is transferred from a cold to a hot reservoir, see Fig. 15.6. As already
known, according to the second law of thermodynamics, work must be supplied. On
the one hand, the benefit of such a process may lie in the heat extracted from the
cold tank. If this is the case, the machine is called a cooling machine. On the other
hand, the purpose may be, for example, to heat a room, so that the benefit is the heat
that is transferred to the hot tank. In such a case, the machine is called a heat pump.
However, a cooling machine and a heat pump have the same technical components
and follow the same thermodynamic principle. The difference between them is only
in the definition of their utility.

22.2.1 Mechanical Compression

To realise a counterclockwise cycle, the fluid must be compressed from a lower


working pressure to a higher working pressure, see Sect. 15.2.1. This can be achieved

6 It is assumed that ea = 0.


7 In this case state (2) shall be in (’) state, see Fig. 22.10b.
702 22 Thermodynamic Cycles with Phase Change

Fig. 22.12 Compression heat pump

Fig. 22.13 Compression heat pump—layout (left), T, s-diagram (right)

with a compressor in case the fluid is in gaseous state. Figure 22.12 shows exemplary
the layout of a heat pump. Heat is transferred from the ambient heat source, which is
at a low temperature level, to the fluid. The fluid is compressed and releases heat at
a higher temperature level. This can be an external water circuit to operate a radiator
for heating purposes. To close the circuit, the fluid must then be expanded to the
lower pressure level. The changes of state are as follows and are illustrated in a
T, s-diagram, see Fig. 22.13:
22.2 Heat Pump and Cooling Machine 703

• (1) → (2): Pressure increase


The fluid is supposed to be superheated, see step (4)–(1). Thus, the required specific
technical work follows

2
wt,12 = wt = v d p + ψ12 > 0. (22.33)
1

Both kinetic and potential energies are ignored as usual. If the compressor is adia-
batic, the indicated increase in specific entropy in the T, s-diagram, see Fig. 22.13,
is due to frictional entropy generation. After the compression the fluid is on a higher
pressure and, according to the first law of thermodynamics, on a higher level of
specific enthalpy, i.e.
wt,12 + q12 = h 2 − h 1 > 0. (22.34)

=0

This increase in enthalpy causes the temperature to rise.


• (2) → (3): Isobaric heat release
Since the fluid is now at a high temperature level, heat can be released to a hot
reservoir, e.g. a room to be heated.8 The heat release causes the fluid to condense
and leave the heat exchanger as a supercooled liquid. However, the heat release
should be isobaric, i.e. if kinetic and potential energies are ignored, there is no
dissipation and thus no entropy generation. The first law of thermodynamics is

q23 = qout = h 3 − h 2 < 0. (22.35)

• (3) → (4): Adiabatic throttling


An adiabatic throttle is applied to reduce the pressure level. As already mentioned,
see Sect. 17.1.4, with a throttle no technical work is released and the pressure drop
is based purely on dissipation, i.e.

4 4
wt,34 = 0 = v d p + ψ34 → v d p = −ψ34 . (22.36)
3 3

Consequently, the specific entropy needs to rise, while the specific enthalpy
remains constant, i.e.
wt,34 + q34 = h 4 − h 3 = 0. (22.37)
 
=0 =0

8 Heating works as long as the fluid temperature is greater than the room temperature.
704 22 Thermodynamic Cycles with Phase Change

• (4) → (1): Isobaric heat supply


To close the thermodynamic cycle, heat must be supplied. As can be seen from
the T, s-diagram, cf. Fig. 22.13, this must be done at a low temperature level
and a correspondingly low pressure level. The thermal energy is to be taken from
the environment, e.g. air, water or soil, which is already at a low temperature
level. For heat transfer, a T > 0 is required, i.e. the temperature of the fluid in
the thermodynamic process must be colder than the ambient temperature by T .
Real fluids, i.e. fluids that can be subject to a phase change, start to vaporise at
low temperatures when the pressure is low. This happens with the change of state
from (4) to (1): Although the ambient temperature is low, heat is supplied to the
even colder fluid in the circuit. The fluid begins to vaporise and if the change of
state is isobaric, the temperature remains constant until the ( ) state is reached. If
the T is still sufficient, the fluid is further superheated until state (1) is finally
reached. The first law of thermodynamics obeys

q41 = qin = h 1 − h 4 > 0. (22.38)

Obviously, according to the T, s-diagram in Fig. 22.13 the process is counterclock-


wise. Since the supply of heat and the release of heat have been assumed to be
isobaric, no entropy is generated. Consequently, both the specific heat supplied qin
and the heat released qout can be identified in the T, s-diagram. According to the
layout of the heat pump, see system boundary in Fig. 22.13, technical work wt is
supplied to the machine. Thus, the overall first law of thermodynamics obeys

|qout | = qin + wt . (22.39)

In HVAC applications, the counterclockwise cycle is often represented in a log p, h-


diagram introduced in Sect. 18.4.3. Such a diagram is exemplary shown in Fig. 22.14
for the discussed cycle. With the premises9 made, the specific heats and the specific
technical works can be easily visualised by horizontal distances, see Fig. 22.14.
Similar to clockwise cycling, the thermal efficiency of counterclockwise cycling can
be calculated by the coefficient of performance, i.e. the ratio of benefit to effort. For
a heat pump it is
|qout | h2 − h3
εHP = = (22.40)
wt,12 h2 − h1

For the cooling machine it is

qin h1 − h4
εCM = = (22.41)
wt,12 h2 − h1

9 Ignoring the kinetic and potential energies.


22.2 Heat Pump and Cooling Machine 705

Fig. 22.14 Compression heat pump—log p, h-diagram

22.2.2 Thermal Compression

In Sect. 22.2.1 it has been shown that the pressure increase is realised with a super-
heated fluid. According to the partial energy equation

2
wt,12 = wt = v d p + ψ12 > 0 (22.42)
1

however, it would be advantageous to increase the pressure of a liquid fluid, as the


specific volume of a liquid is smaller than the specific volume of a gaseous fluid. The
goal of an absorption heat pump is therefore to replace the mechanical compression,
see Fig. 22.15. Apart from the compressor, the components are the same as for
mechanical compression according to Fig. 22.13. The condenser is thus still used
to release heat at a high temperature level, the vaporiser receives the heat at a low
temperature level. The pressure is reduced via a throttle. The components highlighted
in grey replace the former mechanical compressor: The superheated vapour enters
the so-called absorber, which is filled with e.g. water as a solvent or absorbent. The
gaseous working fluid, also called refrigerant, is then absorbed, i.e. the working fluid
forms a liquid mixture with the absorbent. In this process, heat is released at a low
temperature level, so that the exergy content is rather low. The major advantage
of this absorption technique compared to compression heat pumps is that a pump
can now be used to reach the higher pressure level of this liquid mixture. Thus the
technical work is much lower than with a gaseous fluid, see Eq. 22.42. Once the
liquid mixture has reached the upper working pressure, it must be separated again,
i.e. the refrigerant must be released. Generally, the boiling of the refrigerant starts
706 22 Thermodynamic Cycles with Phase Change

Fig. 22.15 Absorption heat pump

earlier than the boiling of the solvent. This principle is utilised in the expeller, where
heat is supplied at a high temperature level so that the refrigerant is separated from
the liquid mixture. The refrigerant in gaseous state leaves the expeller at the high
pressure level and the cycle can continue as before with mechanical compression.
The remaining solvent in the expeller can be throttled and returned to the absorber,
see Fig. 22.15. The technical effort of an absorption heat pump is therefore not
mechanical work, but thermal energy. For this reason, the components outlined in
grey in Fig. 22.15 are referred to as thermal compressor. Figure 22.16 shows an
example of so-called solar cooling. As mentioned before, cooling machine and heat
pump follow the same technical principles. The counterclockwise absorption-based
cycle requires high temperature thermal energy as a driver. If solar energy is available,
this technique can be utilised for such a cooling machine.

Problem 22.1 The high pressure (HP) turbine of a steam power plant expands water
from state (1) with ϑ1 = 540 ◦ C and p1 = 170 bar to state (2) with ϑ2 = 320 ◦ C and
p2 = 35 bar. The fluid is reheated to ϑ3 = 540 ◦ C and finally expands in the low
pressure (LP) turbine to p4 = 0.1 bar. The vapour ratio is x4 = 0.96. Due to friction,
the pressure drop in the reheater is p = 2.5 bar. All changes of state are polytropic
with a constant polytropic exponent. Changes in kinetic and potential energy can be
neglected. Both turbines are adiabatic. The layout of the plant is shown in Fig. 22.17.
(a) Sketch the process qualitatively in a h, s-diagram.
(b) Calculate the specific technical work in both turbines as well as the transferred
specific heat in the reheater.
22.2 Heat Pump and Cooling Machine 707

Fig. 22.16 Solar cooling

Fig. 22.17 Sketch to Problem 22.1

(c) What is the polytropic exponent n in the reheater?


(d) Calculate the specific dissipation in the reheater.
(e) What is the specific exergy loss of the entire plant if the heat supply is realised
with a reservoir of ϑR = 650 ◦ C? Ambient temperature is ϑenv = 17 ◦ C.

Solution

(a) Figure 22.18 shows the process in a h, s-diagram.


(b) To calculate the specific technical work of the turbines, the first law of thermo-
dynamics is applied. For the high pressure turbine it reads as
708 22 Thermodynamic Cycles with Phase Change

Fig. 22.18 h, s-diagram to Problem 22.1

q12 +wt,12 = h 2 − h 1 . (22.43)



=0

The specific enthalpy can be taken from the steam table, i.e.

kJ
h 1 = h ( p1 , ϑ1 ) = 3400.9 . (22.44)
kg

and
kJ
h 2 = h ( p2 , ϑ2 ) = 3030.5 . (22.45)
kg

Hence, the specific technical work of the HP-turbine is

kJ
wt,12 = h 2 − h 1 = −370.4 (22.46)
kg

For the LP turbine, the following results accordingly

q34 +wt,34 = h 4 − h 3 (22.47)



=0

with
kJ
h 3 = h ( p3 , ϑ3 ) = 3544.6 . (22.48)
kg
22.2 Heat Pump and Cooling Machine 709

State (4) follows

kJ
h 4 = h 4 · (1 − x4 ) + h 4 x4 = 2488.2 (22.49)
kg

with
kJ
h 4 ( p4 ) = 191.8123 (22.50)
kg

and
kJ
h 4 ( p4 ) = 2583.9 . (22.51)
kg

Hence, the specific technical work of the LP-turbine is

kJ
wt,34 = h 4 − h 3 = −1056.4 (22.52)
kg

The specific heat in the reheater obeys the first law of thermodynamics

kJ
q23 + wt,23 = h 3 − h 2 = 514.14 (22.53)
 kg
=0

(c) A polytropic change of state is defined by

pvn = const. (22.54)

Thus, it is
p2 v2n = p3 v3n . (22.55)

Solving for n yields


p2
log p3
n=
= 0.1635 (22.56)
v3
log v2

with
m3
v3 = 0.1131 (22.57)
kg

and
m3
v2 = 0.0719 . (22.58)
kg

Both specific volumes have been taken from the steam table. Note that
710 22 Thermodynamic Cycles with Phase Change

Fig. 22.19 Entropy balance


to Problem 22.1

n−1
T3 p3 n
= (22.59)
T2 p2

cannot be applied because the vapour does not obey the ideal gas law, see
Sect. 13.6.
(d) To calculate the specific dissipation in the reheater, the partial energy equation
is applied, i.e.
3
wt,23 = 0 = v d p + ψ23 + ea,23 . (22.60)
  
2 =0

Solving for the specific dissipation10 leads to

3 n−1 
n p3 n kJ
ψ23 = − v d p = −v2 p2 = 22.67 (22.61)
n−1 p2 kg
2

(e) The specific loss of exergy of the entire plant follows

ex,V = Tenv si,14 . (22.62)

Therefore, an entropy balance according to Fig. 22.19 has to be carried out. In


steady state, the incoming entropy must be balanced by the outgoing entropy.
Hence, the second law of thermodynamics obeys

ṁs1 + Ṡi,14 + Ṡa = ṁs4 (22.63)

so that the specific generation of entropy is

10 See Fig. 13.13.


22.2 Heat Pump and Cooling Machine 711

si,14 = s4 − s1 − sa,23 . (22.64)

For the specific entropy that is carried along with the heat, it is11 :

q23 kJ
sa,23 = = 0.5569 . (22.65)
TR kg K

Note that the specific entropy generation si,14 includes the specific entropy gen-
eration due to dissipation in both turbines, the reheater and the imperfection of
the heat transfer.12 The specific entropy of state (1) can be taken out of the steam
table, i.e.
kJ
s1 = s ( p1 , ϑ1 ) = 6.4106 . (22.66)
kg K

For state (4) it is


kJ
s4 = s4 · (1 − x4 ) + s4 x4 = 7.8489 (22.67)
kg K

with kJ
s4 ( p4 ) = 0.6492 (22.68)
kg K

and kJ
s4 ( p4 ) = 8.1489 . (22.69)
kg K

Hence, the specific generation of entropy is


kJ
si,14 = 0.8813 . (22.70)
kg K

The specific loss of exergy finally is

kJ
ex,V = Tenv si,14 = 255.7 (22.71)
kg

Problem 22.2 A compression refrigerator operated with ammonia as working fluid


is to maintain a room at a temperature of ϑ0 = −12 ◦ C. For this purpose, the cooling
machine is supplied with a cooling power of Q̇ = 135 kW. The vaporisation pressure
is p0 = 2.264 bar. The vapour in the saturated state ( ) is compressed in an adiabatic
compressor (ηs,V = 0.785) to a pressure of p = 11.666 bar. After complete isobaric
condensation, i.e. to state ( ), the fluid is adiabatically throttled to a pressure of p0 . The
condenser is operated with cooling water entering at a temperature of ϑenv = 15 ◦ C.

11 This is because the temperature of the heat reservoir is constant.


12 This imperfection is due to the T between fluid and reservoir.
712 22 Thermodynamic Cycles with Phase Change

The cooling water shall be treated as an incompressible liquid. There shall be no


pressure drop for the cooling water. Changes in kinetic as well as potential energy
are to be neglected.
(a) Sketch the layout of the cooling machine and illustrate the changes of state
qualitatively in a T, s-diagram. Additionally mark the temperatures T0 and Tenv .
Clearly identify the states (1)–(4).
(b) Calculate the mass flux of the ammonia ṁ A .
(c) What is the technical power of the compressor? Determine the coefficient of
performance of the cooling machine ε. What is the benchmarking coefficient of
performance?
(d) The maximum temperature of the cooling water shall not exceed ϑW = 27 ◦ C.
What is the required mass flux of the cooling water ṁ W ? The specific heat
capacity of the water is cw = 4.187 kgkJK .
(e) What is the flux of exergy loss  Ė x,V in the condenser?
For ammonia in saturated state it is

p[bar] ϑs [◦ C] kJ
h[ kg ] s[ kgkJK ]
2.264 –16.0 h  = 288.5, h  = 1604.1 s  = 1.2877, s  = 6.4038
11.666 30.0 h  = 503.6, h  = 1648.1 s  = 2.0512, s  = 5.8267

For ammonia at p = 11.666 bar it is

ϑ[◦ C] kJ
h[ kg ] s[ kgkJK ]
100 1840.1 6.3988
110 1865.2 6.4653
120 1890.1 6.5297
130 1914.9 6.5919

Solution

(a) The layout of this process is shown in Fig. 22.20a. Figure 22.20b illustrates
the changes of state in a T, s-diagram. Note that the ambient temperature Tenv
must be lower than the temperatures T2 and T3 , respectively, because the fluid
has to release heat to the environment. T0 must be correspondingly greater than
T1 = T4 in order to supply heat to the fluid.
(b) The mass flux ṁ A is obtained by applying the first law of thermodynamics for
the vaporiser, i.e.
Q̇ = +135 kW = ṁ A (h 1 − h 4 ) . (22.72)

The throttling is adiabatic and kinetic respectively potential energies are negli-
gible, so that the change of state is isenthalpic, i.e.
22.2 Heat Pump and Cooling Machine 713

Fig. 22.20 a Sketch of the plant, b T, s-diagram to Problem 22.2

kJ
h 4 = h 3 = h  = 503.6 . (22.73)
kg

For state (1) it is known, that

kJ
h 1 = h  = 1604.1 . (22.74)
kg

Thus, the mass flux follows

Q̇ kg
ṁ A = = 0.1227 (22.75)
h1 − h3 s

(c) The technical power of the compressor results from the first law of thermody-
namics, i.e.
Pt,12 = ṁ A (h 2 − h 1 ) . (22.76)

By introducing the isentropic efficiency ηs,V it results in

(h 2s − h 1 )
Pt,12 = ṁ A . (22.77)
ηs,V

For state (1) it is known, that


714 22 Thermodynamic Cycles with Phase Change

kJ
h 1 = h  = 1604.1 (22.78)
kg

and
kJ
s1 = s  = 6.4038 . (22.79)
kg K

The hypothetical change of state from (1) to (2s) is isentropic, i.e.

kJ
s2s = s1 = 6.4038 . (22.80)
kg K

Together with the pressure p2s = p2 = 11.666 bar, the specific enthalpy is
obtained by linear interpolation from the vapour data, i.e.

kJ
h 2s = 1842.0 . (22.81)
kg

Finally, the technical power of the compressor obeys

(h 2s − h 1 )
Pt,12 = ṁ A = 37.17 kW (22.82)
ηs,V

The specific enthalpy of state (2) then is


Pt,12 kJ
h2 = h1 = = 1907.1 . (22.83)
ṁ A kg
The coefficient of performance for the cooling machine is


ε= = 3.6313 (22.84)
Pt,12

A perfect process, i.e. a Carnot process, would result in an efficiency of

Tmin
εC = = 9.6722 (22.85)
Tmax − Tmin

(d) The mass flux for the cooling water ṁ W is obtained by applying the first law of
thermodynamics for the condenser, cf. Fig. 22.21a, i.e.

ṁ A h 2 + ṁ W h env = ṁ A h 3 + ṁ W h W . (22.86)

Rearrangement results in
22.2 Heat Pump and Cooling Machine 715

Fig. 22.21 a Energy balance, b Entropy balance for the condenser

ṁ A h 2 − ṁ A h 3 = ṁ W h W − ṁ W h env . (22.87)

Applying the caloric equation of state for the cooling water13

ṁ A h 2 − ṁ A h 3 = ṁ W cW (ϑW − ϑenv ) . (22.88)

Thus, the mass flow rate of the cooling water is

ṁ A (h 2 − h 3 ) kg
ṁ W = = 3.4268 (22.89)
cW (ϑW − ϑenv ) s

(e) The loss of exergy  Ė x,V in the condenser follows

 Ė x,V = Tenv Ṡi . (22.90)

Therefore, an entropy balance is required to calculate entropy generation, see


Fig. 22.21b. This balance leads to

ṁ A s2 + ṁ W senv + Ṡi = ṁ A s3 + ṁ W sW . (22.91)

Hence, the entropy generation obeys

Ṡi = ṁ A (s3 − s2 ) + ṁ W (sW − senv ) . (22.92)

The application of the caloric equation of state for the cooling water, which is
assumed to be incompressible, leads to
TW
Ṡi = ṁ A (s3 − s2 ) + ṁ W cW ln . (22.93)
Tenv
The specific entropy of state (3) is

13Cooling water is regarded as an incompressible liquid. There shall be no pressure drop for the
cooling water.
716 22 Thermodynamic Cycles with Phase Change

kJ
s3 = s  = 2.0512 . (22.94)
kg K

For state (2) it follows

kJ
s2 = s (h 2 , p2 ) = 6.5724 . (22.95)
kg K

Thus, the generation of entropy14 is

TW kW
Ṡi = ṁ A (s3 − s2 ) + ṁ W cW ln = 0.0308 . (22.96)
Tenv K

Finally, the loss of exergy is

 Ė x,V = Tenv Ṡi = 8.8701 kW (22.97)

14 The generation of entropy is due to the heat transfer from ammonia to the cooling water. Since
there is no pressure drop for the ammonia or the cooling water, there is also no dissipation.
Part III
Reactive Systems
Chapter 23
Combustion Processes

In Part I, ideal gases and incompressible liquids have been introduced and the basic
thermodynamic correlations have been derived. Part II has shown that real fluids can
be subject to a change of aggregate state and that many thermodynamic cycles are
based on phase change, e.g. a steam power plant. Furthermore, Part II included mix-
ture of fluids, e.g. humid air or mixtures of ideal gases. However, in these mixtures,
each component is stable and not part of a chemical reaction in which one or more
chemical compounds are converted into others. In Part III, the focus is now on chemi-
cally reacting systems: First, the stoichiometry of a chemical reaction is investigated,
i.e. the principle of conservation of substances is applied to reactants and products of
a chemical reaction. By doing so, it is possible to predict the composition of the prod-
ucts if it is known in which ratio the reactants react. This is important, for example,
when the composition of the exhaust gas of a combustion process must comply with
technical threshold values. Secondly, chemically reacting systems are investigated
with regard to their energy balance. In this chapter, the conventional heating value
approach is followed, focusing on technical combustion, i.e. combustion based on
fossil fuels. In Chap. 24, a different energetic approach based on absolute enthalpy
respectively entropy is presented. The major advantage of this method is that the
irreversibility of chemical reactions can be quantified. Based on the irreversibilities
and the associated entropy generation, an exergetic evaluation can also be carried
out. This is not possible with the heating value method.

23.1 Fossil Fuels

A combustion is a chemical reaction of fuels with oxygen, i.e. an oxidation. The


oxygen is usually taken from the atmospheric air. If the reaction is exothermic,
heat is released that can be used in a steam power process, for example, cf. Fig. 23.2.
Obviously, the chemically bonded energy of the fuel is converted into thermal energy.
© Springer Nature Switzerland AG 2022 719
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_23
720 23 Combustion Processes

Fig. 23.1 Fossil fuels according to [6]

As already mentioned, the focus in this chapter is on fossil fuels. Fossil fuels are
primarily composed of carbon, hydrogen and oxygen. In addition, sulphur can also
be converted by oxidation:
• Carbon C → CO2
• Hydrogen1 H2 → H2 O
• Sulphur S → SO2
In general, fuels consist of the listed combustible components (C, H2 , S) and non-
combustible components, e.g. O2 , N2 , CO2 , H2 O, noble gases and ashes. For solid
fossil fuels, the ratio of carbon and hydrogen to oxygen content varies with the age
of the fuel, see Fig. 23.1. This degree of carbonisation is much higher for coke, for
example, than for lignite. However, fossil fuels contain many other components, e.g.
water, ash and nitrogen N2 , but are treated as inert, i.e. not chemically reactive, in
this context. Nevertheless, they have a energetic influence on the chemical reaction.
In the case of nitrogen, the reactive part is much more complex and subject of many
other books, see [6, 41]. If the oxygen supply is too low, oxidation is incomplete, i.e.
the exhaust gas still contains reactants that could be further converted. This can be
the fossil fuel itself or not yet fully oxidised products such as carbon monoxide CO,
which could be further oxidised if oxygen were available. In this chapter, the focus
is on stoichiometric reactions with sufficient oxygen for a complete conversion of
the fuel components carbon, hydrogen and sulphur.

1 Hydrogen in its stable form under atmospheric conditions is molecular.


23.1 Fossil Fuels 721

However, fossil fuels can occur in different aggregate states under standard con-
ditions, i.e.
• Solid fuels:
Hard coal, lignite, biomass
• Liquid fuels:
Oil, diesel, hydrocarbons
• Gaseous fuels:
Hydrogen H2
Carbon monoxide CO
Hydrocarbons: methane CH4 , natural gas, biogas, acetylene C2 H2 , ethylene C2 H4 ,
ethane C2 H6 or others Cx H y
In order to conduct an oxidation, the ignition temperature, i.e. the temperature at
which the reaction continues by itself under heat release, must be reached. In addition,
a sufficient amount of oxygen must be present. For gaseous fuels, the mixture of fuel
and oxygen must be within the ignition limit, as only then is the released heat sufficient
to keep the temperature above the ignition temperature. The ignition limits depend
on pressure and temperature. At standard pressure p0 = 1 bar and temperature ϑ0 =
20 ◦ C, these ignition limits2 for natural gas-air mixtures are approx. 4.216.5, for
hydrogen-air mixtures about 477, see [42]. During the combustion of liquid or solid
fuel, the non-visible volatile components are burnt first. As the temperature in the
combustion chamber rises, the solid/liquid components of the fuel are also split into
combustible gases and can react with the oxygen in the combustion air.
Figure 23.2 shows a combustion process in black box notation. A fuel mass
flow and an air/oxygen mass flow are supplied to the combustion chamber. The
exhaust gas, i.e. the product of the chemical reaction, leaves the system boundary.
The released heat can be used, for example, to operate a steam power plant. To
calculate such a process, two steps are carried out in this chapter:
• Step 1: Stoichiometry

Substance balances can be carried out on the basis of the chemical reaction scheme.
It is clarified which amount of oxygen is needed for a complete oxidation of the
fuel. In addition, the composition of the exhaust gas is calculated. To do so, the
composition of the reactants must first be determined.

• Step 2: Energetic balancing

Similar to the non-reacting open systems in Parts I and II, an energy balance is
performed. The energetic aspect of the material conversion, i.e. the oxidation of
the fuel, must be taken into account. Such a balance allows the calculation of the
exhaust gas temperature, respectively the amount of heat released. Furthermore, it
can be predicted, for example, whether water will condense from the exhaust gas.

VF
2 The ignition limits are defined as σ = Vtotal .
722 23 Combustion Processes

Fig. 23.2 Combustion process

23.2 Fuel Composition

In order to carry out mass and energy balances of the combustion process, the fuel
composition is needed first. However, a distinction is made according to the aggregate
state of the fuel.

23.2.1 Solid Fuels

The composition of solid and liquid fuels is usually given by mass fractions in relation
to the mass of the fuel m F , i.e.
• Carbon mC
ξC = =c (23.1)
mF

• Hydrogen
m H2
ξH2 = =h (23.2)
mF

• Sulphur
mS
ξS = =s (23.3)
mF

• Oxygen
m O2
ξO2 = =o (23.4)
mF
23.2 Fuel Composition 723

• Nitrogen
m N2
ξN2 = =n (23.5)
mF

• Ashes m Ash
ξAsh = =a (23.6)
mF

• Water mW
ξW = =w (23.7)
mF

The total mass of the fuel m F is

m F = m C + m H2 + m S + m O2 + m N2 + m Ash + m W . (23.8)

A division by m F yields

c+h+s+o+n+a+w =1 (23.9)

23.2.2 Liquid Fuels

Liquid fuels consist mainly of hydrocarbons. They contain carbon C, hydrogen H2 ,


oxygen O2 and sulphur S. Their composition is usually also expressed in mass frac-
tions, i.e.
• Carbon mC
ξC = =c (23.10)
mF

• Hydrogen
m H2
ξH2 = =h (23.11)
mF

• Sulphur
mS
ξS = =s (23.12)
mF

• Oxygen
m O2
ξO2 = =o (23.13)
mF

Due to
m F = m C + m H2 + m S + m O2 (23.14)

it is
c+h+s+o=1 (23.15)
724 23 Combustion Processes

23.2.3 Gaseous Fuels

Gaseous fuels are usually gas mixtures whose components must be determined by
elemental analysis. The concentration of a component k of the mixture is given in
volume fractions related to the total volume of the fuel VF , i.e.

Vk
σk = . (23.16)
VF

If the components can be treated as ideal gases, the volume fraction in a gas mixture
is equal to its molar fraction, see Part II, Sect. 19.3.1, i.e.
nk
σk = xk = . (23.17)
nF

23.3 Stoichiometry

The following chemical equations show the relevant chemical reactions of the com-
bustible components,3 i.e. C, H, S, of fossil fuels and oxygen under stoichiometric4
conditions, i.e.

• Carbon:
C + O2 → CO2 (23.18)

• Hydrogen:
1
H2 + O2 → H2 O (23.19)
2
• Sulphur:
S + O2 → SO2 (23.20)

Chemical equations are not quantity equations in the mathematical sense, but stoi-
chiometric equations, i.e. they follow the law of conservation of number of atoms:
The number of atoms of each element on the reactant side must be equal to the num-
ber of atoms of the same element on the product side. For example, Eq. 23.18 has
one C atom on the left side and one C atom on the right side. The number of O atoms
is two on the left and two on the right. Consequently, the mass conservation of Eq.
23.18 is fulfilled. Actually, stoichiometry shows the numerical ratio of the elements
within the reaction.

3 The other components such as nitrogen, water and ashes are treated as inert, i.e. not chemically
reactive.
4 Each combustible element is therefore completely oxidised. The oxygen on the left side of the

equation is completely consumed, so that no residual oxygen remains on the right side.
23.3 Stoichiometry 725

Fig. 23.3 Example of a solid fossil fuel

23.3.1 Solid/Liquid Fuels

Air/Fuel Composition

Oxygen Demand
In this section, the minimum oxygen demand is investigated, i.e. the chemical Eqs.
23.18–23.20 are applied to clarify how much oxygen is required to completely oxidise
all combustible components of a solid/liquid fossil fuel. The composition of this fuel
is shown in Fig. 23.3. Note that the combustible elements are marked by a box. To
determine the minimum oxygen demand, each combustible element is analysed step
by step, starting with carbon:
• Step 1 (Carbon): C + O2 → CO2

The chemical reaction obeys

1 atom C + 1 molecule O2 → 1 molecule CO2 . (23.21)

Multiplied with the Avogadro constant, i.e. 6.022 × 1026 particles represent
1 kmol:
1 kmol C + 1 kmol O2 → 1 kmol CO2 . (23.22)

By introducing the molar masses,5 it finally is

12 kg C + 32 kg O2 → 44 kg CO2 . (23.23)

5 1 kmol of C has a mass of 12 kg. 1 kmol of O has a mass of 16 kg. 1 kmol of CO2 has a mass of
1 · 12 kg + 2 · 16 kg = 44 kg. See also Table 3.2.
726 23 Combustion Processes

According to the rule of three, this equation can be given in a general notation, i.e.
divide by 12 kg and multiply by the mass of carbon in the fuel m C , i.e.

m C (C) + 2.667 · m C (O2 ) → 3.667 · m C (CO2 ) (23.24)

This equation reads like a recipe: For complete oxidation, 2.667 times the mass
of present carbon is required as the mass of oxygen. Oxygen and carbon lead to
3.667 times the mass of carbon as the mass of carbon dioxide. In other words,
1 kg carbon requires, for example, 2.667 kg oxygen and reacts to 3.667 kg carbon
dioxide.

• Step 2 (Hydrogen): 2H2 + O2 → 2H2 O

The chemical reaction reads as

2 molecules H2 + 1 molecule O2 → 2 molecules H2 O. (23.25)

Analogous to step 1, multiplication by the Avogadro constant yields

2 kmol H2 + 1 kmol O2 → 2 kmol H2 O. (23.26)

And the multiplication with the molar masses leads to the mass balance

4 kg H2 + 32 kg O2 → 36 kg H2 O. (23.27)

A generalisation based on the rule of three finally yields the equation

m H2 (H2 ) + 8 · m H2 (O2 ) → 9 · m H2 (H2 O) (23.28)

Complete oxidation requires 8 times the mass of available hydrogen as the mass
of oxygen. Oxygen and hydrogen lead to 9 times the mass of hydrogen as the mass
of water.

• Step 3 (Sulphur): S + O2 → SO2

The chemical reaction obeys

1 atom S + 1 molecule O2 → 1 molecule SO2 . (23.29)

Multiplication by the Avogadro constant results in

1 kmol S + 1 kmol O2 → 1 kmol SO2 (23.30)

and multiplying by the molar masses yields

32 kg S + 32 kg O2 → 64 kg SO2 . (23.31)
23.3 Stoichiometry 727

Again, a generalisation based on the rule of three eventually leads to the equation

m S (S) + m S (O2 ) → 2 · m S (SO2 ) (23.32)

For complete oxidation, the mass of available sulphur is needed as the mass of
oxygen. Oxygen and sulphur lead to a doubling of the mass of sulphur as the mass
of sulphur dioxide.
In steps 1–3, all combustible elements have now been oxidised. A combination of
Eqs. 23.24, 23.28 and 23.32 gives the minimum oxygen demand. The amount of
oxygen to be supplied is reduced by the mass of oxygen m O2 , which may be bound
in the fuel itself, see Fig. 23.3. Thus, the minimum oxygen demand m O2 ,min obeys

m O2 ,min = 2.667 · m C + 8 · m H2 + m S − m O2 . (23.33)

Division by the mass of the fuel m F as a reference quantity leads to


m O2 ,min mC m H2 mS m O2
= 2.667 · +8· + − . (23.34)
mF mF mF mF mF

Applying the abbreviations already introduced, see Sect. 23.2, the mass-specific
minimum oxygen demand results in

m O2 ,min
omin ≡ = 2.667 · c + 8 · h + s − o (23.35)
mF

Air Demand
Usually, the reaction is not performed with pure oxygen, but with atmospheric air.
Dry, atmospheric air consists mainly of nitrogen N2 and oxygen O2 , see Chap. 19. Its
composition is now simplified to consist only of these two components, i.e. carbon
dioxide and argon, for example, are ignored. Hence, in the following Table 23.1,
both the mass fractions and the molar fractions are summarised. A conversion from
molar to mass fractions has been introduced in Chap. 19 and follows, see Eq. 19.64,

Mi
ξi = xi . (23.36)
Mtotal

The minimum air demand, related to the mass of the fuel, can therefore be calcu-
lated as follows

Table 23.1 Simplified composition of dry, atmospheric air


N2 O2
Mass fraction ξi 0.77 0.23
Molar fraction xi 0.79 0.21
728 23 Combustion Processes

m a,min m a,min m O2 ,min 1


lmin = = · = · omin (23.37)
mF m O2 ,min mF ξO2 ,dry air

This equation states that the mass of air is greater than the mass of oxygen, because
air is not only composed of oxygen but also contains a large amount of nitrogen. If
humid air is used instead of dry air, the amount of air required increases further due
to the water contained in the air, i.e.

m A,min = m a,min + m w = m a,min + x · m a,min = m a,min (1 + x). (23.38)


   
dry air water

Note that x is the water content of air, cf. Chap. 20.


Air-Fuel Equivalence Ratio
So far, the minimum oxygen/air demand has been derived. To ensure that oxidation
is complete, the amount of air is often increased further. Therefore, the air-fuel
equivalence ratio λ is defined. This is the ratio of the supplied air mass to the minimum
air mass required for stoichiometric combustion, i.e. complete oxidation:

ma l
λ= = (23.39)
m a,min lmin

However, the range of λ is defined as follows



⎨ < 1, lack of air, rich mixture
λ = 1, stoichiometric combustion (23.40)

> 1, excess air, lean mixture.

Diesel-engines e.g. are operated with λ = 1.1 . . . 1.15, gasoline-engines with cat-
alytic converter with λ = 0.99 . . . 1.00. Gas turbines are operated with λ > 5, in
order to reduce the flame temperature, see Sect. 23.4. If humid air is used instead of
dry air, the required mass of air must be corrected by (1 + x), since

m A = m a + m w = m a + x · m a = m a (1 + x) = λ · m a,min (1 + x). (23.41)

Rich mixtures are not covered in this book because the mechanisms of incomplete
oxidation are rather complex. However, the subject is treated in detail in other text-
books, see [6, 43]. Therefore, the focus here is on oxidations with λ ≥ 1.

Exhaust Gas Composition

Now that it is known how air and fuel must be mixed to achieve complete oxidation,
the next step is to examine the exhaust gas composition. It is assumed that the
23.3 Stoichiometry 729

combustion air consists only of nitrogen and oxygen. According to the derived Eqs.
23.24, 23.28 and 23.32 for the combustible elements, the composition can be derived.
Steps 1–3 have shown, that

m C (C) + 2.667 · m C (O2 ) → 3.667 · m C (CO2 ). (23.42)

and
m H2 (H2 ) + 8 · m H2 (O2 ) → 9 · m H2 (H2 O) (23.43)

and
m S (S) + m S (O2 ) → 2 · m S (SO2 ). (23.44)

According to these equations, the following results for the exhaust gas:
• Carbon dioxide CO2
m EG,CO2 = 3.667 · m C (23.45)

This mass can be referred to the mass of the fuel to obtain the mass-specific exhaust
gas composition of carbon dioxide, i.e.

m EG,CO2 mC
μEG,CO2 = = 3.667 · = 3.667 · c (23.46)
mF mF

• Sulphur dioxide SO2


m EG,SO2 = 2 · m S (23.47)

This mass can be referred to the mass of the fuel to obtain the mass-specific exhaust
gas composition of sulphur dioxide, i.e.

m EG,SO2 mS
μEG,SO2 = =2· =2·s (23.48)
mF mF

• Water H2 O

There are several reasons for the presence of water in the exhaust gas. Due to
the hydrogen content of fossil fuels, water is formed by oxidation as described in
Eq. 23.28. However, fossil fuels can also contain water, see Fig. 23.3, which is
inert but also leaves the combustion process and is part of the exhaust gas. Finally,
the combustion can be operated with humid air, so that the water carried by the air
additionally leaves with the exhaust gas without reacting chemically, i.e.

m EG,H2 O = 9 · m H2 + m W + λ · m a,min · x . (23.49)


   
fuel water water from the comb. air
730 23 Combustion Processes

This mass can be referred to the mass of the fuel, so that the mass-specific exhaust
gas composition of water is obtained, i.e.

m EG,H2 O
μEG,H2 O = = 9 · h + w + λ · lmin · x (23.50)
mF

• Nitrogen (inert) N2

Actually, there are two sources for the nitrogen: Firstly, the exhaust gas nitrogen
comes from the nitrogen that is part of the fossil fuel, see Fig. 23.3. Secondly, it
comes from the air required for oxidation, i.e.

m EG,N2 = mN + λ · m a,min · ξN2 ,dry air . (23.51)


2   
fuel nitrogen nitrogen from the comb. air

This mass can be referred to the mass of the fuel, so that the mass-specific exhaust
gas composition of nitrogen is obtained, i.e.

m EG,N2
μEG,N2 = = n + λ · lmin · ξN2 ,dry air (23.52)
mF

• Oxygen O2

The reason for the presence of oxygen in the exhaust gas is that the combustion
is carried out with a lean mixture, i.e. λ > 1. The oxygen in the exhaust gas is
the difference between the total oxygen supplied on the reactant side minus the
minimum oxygen demand required for combustion, i.e.

m EG,O2 = λ · m a,min · ξO2 ,dry air − m a,min · ξO2 ,dry air


     
oxygen from the comb. air oxygen req. for comb. (23.53)
= m a,min · ξO2 ,dry air (λ − 1).

This mass can be referred to the mass of the fuel, so that the mass-specific exhaust
gas composition of oxygen is obtained, i.e.

m EG,O2
μEG,O2 = = lmin · ξO2 ,dry air · (λ − 1) = omin · (λ − 1) (23.54)
mF

Apparently, this equation shows that no exhaust gas oxygen is present if the com-
bustion is stoichiometric, i.e. λ = 1.
If the combustion air does not consist simplified only of nitrogen and oxygen, the
exhaust gas composition also contains other components, i.e. the calculation must
be adapted.
23.3 Stoichiometry 731

Exhaust Gas Concentration

In the previous section, the exhaust gas composition μEG,i of a component has been
derived. As it relates to the mass m F of the fuel, it is possible to achieve μEG,i >
1. Thus, the exhaust gas composition is not to be confused with an exhaust gas
concentration. Such a concentration is necessary to determine the caloric and thermal
properties of the exhaust gas, which can be treated, for example, as a mixture of ideal
gases, see Chap. 19. The mass-fraction of a exhaust gas component i is referred to
the mass of the exhaust gas and thus follows
m EG,i m F · μEG,i μEG,i
ξEG,i = = = . (23.55)
m EG m F · μEG μEG

Note that the total mass of the exhaust gas is the sum of the individual masses of its
components, i.e.
m EG = m EG,i . (23.56)
i

Dividing by the mass of the fuel, yields



μEG = μEG,i . (23.57)
i

This simplifies the concentration of the components to


μEG,i
ξEG,i =
. (23.58)
μEG,i
i

Applied to the concentrations of the exemplary exhaust gas, it is therefore:


• Carbon dioxide CO2
m EG,CO2 m F · μEG,CO2 μEG,CO2
ξEG,CO2 = =
=
(23.59)
m EG m F · μEG,i μEG,i
i i

• Sulphur dioxide SO2

m EG,SO2 m F · μEG,SO2 μEG,SO2


ξEG,SO2 = =
=
(23.60)
m EG m F · μEG,i μEG,i
i i

• Water H2 O
m EG,H2 O m F · μEG,H2 O μEG,H O
ξEG,H2 O = =
=
2 (23.61)
m EG m F · μEG,i μEG,i
i i
732 23 Combustion Processes

• Oxygen O2
m EG,O2 m F · μEG,O2 μEG,O2
ξEG,O2 = =
=
(23.62)
m EG m F · μEG,i μEG,i
i i

• Nitrogen N2
m EG,N2 m F · μEG,N2 μEG,N2
ξEG,N2 = =
=
(23.63)
m EG m F · μEG,i μEG,i
i i

Once the concentrations of the individual components in the exhaust gas are known,
the exhaust gas properties can be determined, i.e. c p,EG , REG and MEG . Thus, all the
principles from Chap. 19 can be applied. Note that mass fractions can be converted
to molar fractions according to Eq. 19.64.

23.3.2 Gaseous Fuels

Air/Fuel Composition

Oxygen Demand
In this section, the minimum oxygen demand is investigated, i.e. the chemical Eqs.
23.18–23.20 are applied to clarify how much oxygen is required to completely oxidise
all combustible components of a gaseous fossil fuel. The procedure is analogous to
the approach followed for solid fuels. The exemplary composition of this gaseous
fuel is shown in Fig. 23.4. Note that the combustible elements in this example are

Fig. 23.4 Exemplary gaseous fuel


23.3 Stoichiometry 733

marked by a box. To determine the minimum oxygen demand, each combustible


element is analysed step by step, starting with hydrogen:
• Step 1 (Hydrogen): n H2

The chemical reaction reads as

2H2 + O2 → 2H2 O. (23.64)

Multiplication with the Avogadro constant yields

2 kmol H2 + 1 kmol O2 → 2 kmol H2 O. (23.65)

A generalisation based on the rule of three finally leads to the following equation

n H2 (H2 ) + 0.5 · n H2 (O2 ) → 1 · n H2 (H2 O) (23.66)

• Step 2 (Carbon dioxide): n CO2 → inert, no further reaction


• Step 3 (Oxygen): n O2 → lowers the required feed of oxygen
• Step 4 (Nitrogen): n N2 → inert, no further reaction
• Step 5 (Carbon monoxide): n CO

The chemical reaction obeys

2CO + O2 → 2CO2 . (23.67)

Multiplication with the Avogadro constant yields

2 kmol CO + 1 kmol O2 → 2 kmol CO2 . (23.68)

A generalisation based on the rule of three finally leads to the following equation

n CO (CO) + 0.5 · n CO (O2 ) → 1 · n CO (CO2 ) (23.69)

• Step 6 (Chemical bonds in the form of Ca Hb Oz N p Sr ):

The gaseous fuel can also consist of several hydrocarbon complexes.6 The aim is
to find a generic chemical reaction scheme that allows each of these complexes to
be treated. Such a generic chemical reaction obeys

b z b p
Ca Hb Oz N p Sr + (a + − + r ) · O2 → a · CO2 + · H2 O + r · SO2 + · N2 .
4 2 2 2
(23.70)

6 Such as methane, propane, butane, ethanol or others.


734 23 Combustion Processes

By adjusting the stoichiometric parameters a, b, z, p, r each component can be


considered with this approach.7 Multiplication with the Avogadro constant yields

b z
− + r ) · kmol O2
1 kmol Ca Hb Oz N p Sr + (a +
4 2 (23.71)
b p
→ a · kmol CO2 + · kmol H2 O + r · kmol SO2 + · kmol N2 .
2 2
A generalisation based on the rule of three finally leads to the following equation

b z
n Ca Hb Oz N p Sr (Ca Hb Oz N p Sr ) + (a + − + r ) · n Ca Hb Oz N p Sr (O2 ) →
4 2
a · n Ca Hb Oz N p Sr (CO2 )+
b
+ · n Ca Hb Oz N p Sr (H2 O)+ (23.72)
2
+ r · n Ca Hb Oz N p Sr (SO2 )+
p
+ · n Ca Hb Oz N p Sr (N2 )
2

Now, each of the exemplary components, see Fig. 23.4, has been investigated. The
following applies to the fuel mixture:

n F = n H2 + n CO2 + n O2 + n N2 + n CO + n Ca Hb Oz N p Sr ,i . (23.73)
i

In order to completely oxidise the exemplary fuel mixture analogous to solid fuel,
the oxygen demand results from the sum of the oxygen demands of the individual
components, see steps 1–6:
bi zi
(ai + − + ri ) · n Ca Hb Oz N p Sr ,i − n O2 .
n O2 ,min = 0.5 · n H2 + 0.5 · n CO +
i
4 2
(23.74)
The right-hand side of the equation represents the molar amounts of the individual
components of the fuel. n O2 is the molar amount of oxygen in the fuel mixture, see
Fig. 23.4, so the total oxygen demand is reduced by this amount. Dividing by the
molar amount of the total fuel mixture n F gives the molar specific minimum oxygen
demand of the fuel, i.e.
n O ,min
Omin = 2 . (23.75)
nF

7Actually, steps 1–5 could be skipped, as the primary reactions can also be dealt with using this
generic approach.
23.3 Stoichiometry 735

Thus, it is

bi zi
Omin = 0.5 · xH2 + 0.5 · xCO + (ai + − + ri ) · xCa Hb Oz N p Sr ,i − xO2
i
4 2

(23.76)

Note that not only one hydrocarbon complex can be part of the gaseous fuel, but
several. Each of them has its own specific oxygen demand to oxidise completely.
Air Demand
Usually, the reaction is not performed with pure oxygen, but with atmospheric air.
Dry, atmospheric air consists mainly of nitrogen N2 and oxygen O2 , see Chap. 19.
In simplified terms, its composition now consists only of these two components, i.e.
without taking into account e.g. carbon dioxide and argon, see e.g. Table 23.1. The
minimum demand of air, referred to the molar quantity of the fuel, can be calculated
as follows

n a,min n a,min n O2 ,min 1


L min = = · = · Omin (23.77)
nF n O2 ,min nF xO2 ,dry air

This equation states that the molar amount of air is greater than the molar amount of
oxygen, because air not only consists of oxygen but also contains a large amount of
nitrogen. If humid air is used instead of dry air, the amount of air required increases
further due to the water contained in the air, i.e.

m A,min = m a,min + m v = m a,min + x · m a,min = m a,min (1 + x). (23.78)

This results in a corresponding correlation for the molar amount:


• Dry air
m a,min = n a,min Ma (23.79)

so, that
m a,min
n a,min = (23.80)
Ma

• Water
m w = n w Mw = x · m a,min = x · n a,min Ma (23.81)

so, that
Ma
n w = x · n a,min = 1.608 · x · n a,min (23.82)
Mw
736 23 Combustion Processes

Ma
Note that the ratio M w
= 1.608 has been derived in Sect. 20.1.4 and only applies
to a standard atmosphere.8
If the combustion air is humid, the minimum molar amount of humid air is therefore

n A,min = n a,min + n w = n a,min (1 + 1.608 · x) (23.83)

Air-Fuel Equivalence Ratio


The air-fuel equivalence ratio λ has been introduced earlier for solid fuels and is now
adapted for gaseous fuels. It is the ratio of the supplied air mass to the minimum air
mass required for stoichiometric combustion, i.e. complete oxidation:

ma n a Ma na L
λ= = = = (23.84)
m a,min n a,min Ma n a,min L min

However, the range of λ is defined as follows



⎨ < 1, lack of air, rich mixture
λ = 1, stoichiometric combustion (23.85)

> 1, excess air, lean mixture.

If humid air is used instead of dry air, the required amount of air must be

n A = λ · n A,min . (23.86)

In combination with Eq. 23.83 it is

n A = λ · n A,min = λ · n a,min (1 + 1.608 · x). (23.87)

Exhaust Gas Composition

Now that it is known how air and fuel must be mixed to achieve complete oxidation,
the next step is to examine the exhaust gas composition. It is assumed that the
combustion air consists only of nitrogen and oxygen. According to the derived Eqs.
23.66–23.72 for the exemplary gaseous fuel, the composition of the exhaust gas can
be determined.
• Carbon dioxide CO2

In this example, the carbon dioxide is a result of the carbon monoxide of the fuel,
which oxidises completely during combustion, and a result of the oxidation of the

8Here the value 1.608 is taken, although the standard atmosphere was reduced to nitrogen and
oxygen. The associated error is to be ignored.
23.3 Stoichiometry 737

hydrocarbon complexes, see Fig. 23.4, i.e.



n EG,CO2 = 1 · n CO + ai · n Ca Hb Oz N p Sr ,i + n CO2 . (23.88)
i

This molar amount can be referred to the molar amount of the fuel to obtain the
molar specific exhaust gas composition of the carbon dioxide, i.e.

n EG,CO2
νEG,CO2 = = xCO + ai · xCa Hb Oz N p Sr ,i + xCO2 (23.89)
nF i

• Sulphur dioxide SO2

The sulphur dioxide in the exhaust gas is due to the oxidation of the hydrocarbon
complexes, i.e.
n EG,SO2 = ri · n Ca Hb Oz N p Sr ,i . (23.90)
i

This molar amount can be referred to the molar amount of the fuel to obtain the
molar specific exhaust gas composition of the sulphur dioxide, i.e.

n EG,SO2
νEG,SO2 = = ri · xCa Hb Oz N p Sr ,i (23.91)
nF i

• Water H2 O

The water is the result of the combustion of the hydrogen of the fuel, the hydrocar-
bon complexes of the fuel, see Fig. 23.4, and the water that is part of the combustion
air, i.e.
bi
n EG,H2 O = n H2 + · n Ca Hb Oz N p Sr ,i + 1.608 · λ · n a,min · x . (23.92)
2   
i
water from the comb. air

This molar amount can be referred to the molar amount of fuel to obtain the molar
specific exhaust gas composition of water, i.e.

n EG,H2 O bi
νEG,H2 O = = x H2 + · xCa Hb Oz N p Sr ,i + 1.608 · λ · L min · x
nF i
2

(23.93)
738 23 Combustion Processes

• Nitrogen N2

The nitrogen is considered inert and is passed through the combustion process.
The source of the exhaust nitrogen is the nitrogen of the fuel, the bound nitrogen
in the hydrocarbon complexes of the fuel, see Fig. 23.4, and the nitrogen that is
part of the oxidation air, i.e.
pi
n EG,N2 = n N2 + · n Ca Hb Oz N p Sr ,i + λ · n a,min · xN2 ,dry air . (23.94)
 2   
i
fuel nitrogen    nitrogen from the comb. air
from the CHONS-complexes

This molar amount can be referred to the molar amount of the fuel to obtain the
molar specific exhaust gas composition of the nitrogen, i.e.

n EG,N2 pi
νEG,N2 = = x N2 + · xCa Hb Oz N p Sr ,i + λ · L min · xN2 ,dry air (23.95)
nF i
2

• Oxygen O2

The reason for the presence of oxygen in the exhaust gas is that the combustion
is carried out with a lean mixture, i.e. λ > 1. The oxygen in the exhaust gas is
the difference between the total oxygen supplied on the reactant side minus the
minimum oxygen demand required for combustion, i.e.

n EG,O2 = λ · n a,min · xO2 ,dry air − n a,min · xO2 ,dry air


     
oxygen from comb. air oxygen req. for comb. (23.96)
= n a,min · xO2 ,dry air (λ − 1).

This molar amount can be referred to the molar amount of fuel to obtain the molar
specific exhaust gas composition of oxygen, i.e.

n EG,O2
νEG,O2 = = L min · xO2 ,dry air · (λ − 1) = Omin · (λ − 1) (23.97)
nF

If the combustion air is not simplified to consist only of nitrogen and oxygen, the
exhaust gas composition also contains other components, i.e. the calculation must be
adjusted. The same applies to gaseous fuels that have a different composition than
shown in Fig. 23.4.

Exhaust Gas Concentration

In the previous section, the exhaust gas composition νEG,i has been derived. Since it
refers to the molar amount of fuel n F , it is possible to achieve νEG,i > 1. The exhaust
gas composition is therefore not to be confused with an exhaust gas concentration.
23.3 Stoichiometry 739

Such a concentration is needed to define the caloric and thermal properties of the
exhaust gas, which can be regarded as a mixture of ideal gases, for example, see
Chap. 19. The molar-fraction of an exhaust gas component i is referred to the molar
quantity of the exhaust gas and thus reads as follows
n EG,i n F · νEG,i νEG,i
xEG,i = = = . (23.98)
n EG n F · νEG νEG

Note that the total molar amount of the exhaust gas is the sum of the individual molar
amounts of its components, i.e.

n EG = n EG,i . (23.99)
i

Dividing by the molar quantity of the fuel, one gets



νEG = νEG,i . (23.100)
i

Consequently, the concentration of the components simplifies to


νEG,i
xEG,i =
. (23.101)
νEG,i
i

Referring to the concentrations of the exemplary exhaust gas, it is therefore:


• Carbon dioxide CO2
n EG,CO2 n F · νEG,CO2 νEG,CO2
xEG,CO2 = =
=
(23.102)
n EG n F · νEG,i νEG,i
i i

• Sulphur dioxide SO2

n EG,SO2 n F · νEG,SO2 νEG,SO2


xEG,SO2 = =
=
(23.103)
n EG n F · νEG,i νEG,i
i i

• Water H2 O
n EG,H2 O n F · νEG,H2 O νEG,H O
xEG,H2 O = =
=
2 (23.104)
n EG n F · νEG,i νEG,i
i i

• Oxygen O2
n EG,O2 n F · νEG,O2 νEG,O
xEG,O2 = =
=
2 (23.105)
n EG n F · νEG,i νEG,i
i i
740 23 Combustion Processes

• Nitrogen N2
n EG,N2 n F · νEG,N2 νEG,N
xEG,N2 = =
=
2 (23.106)
n EG n F · νEG,i νEG,i
i i

Once the concentrations of the individual components in the exhaust gas are known,
the exhaust gas properties can be determined, i.e. c p,EG , REG and MEG . Thus, all prin-
ciples following Chap. 19 can be applied. Note that molar fractions can be converted
to mass fractions according to Eq. 19.64.

23.3.3 Mass Conservation

The conservation of mass must, of course, be fulfilled and obeys in the steady state

m F + λ · m a,min · (1 + x) = m F · μEG,i + m A . (23.107)
    
i
fuel comb. air    fuel ashes
exhaust gas

The left side of this equation summarises the flows in. They must be balanced by the
flows out, which are given on the right-hand side of this equation. It should be noted
that the ash is typically not part of the gaseous exhaust gas, but is treated separately
as an outgoing solid or slag. Dividing by the mass of the fuel m F , one obtains

1 + λ · lmin · (1 + x) = μEG,i + a. (23.108)
i

Rearrangement results in

μEG,i = (1 − a) + λ · lmin · (1 + x) (23.109)
i

Any stoichiometric calculation should be carefully checked for conservation of mass.

23.3.4 Conversions

The major difference between the treatment of solid/liquid and gaseous fuels is obvi-
ously that solid/liquid fuels are mostly treated in mass-specific notation, whereas
gaseous fuels are treated in molar-specific notation. With the knowledge from
Chap. 19, a conversion between these two approaches is simple:
• Molar fraction → mass fraction
23.3 Stoichiometry 741

MEG,i
ξEG,i = xEG,i · (23.110)
MEG

• air demand (mass) → air demand (molar amount)

n air m air MF MF
L= = =l· (23.111)
nF m F Mair Mair

• Exhaust gas composition (molar amount) → Exhaust gas composition (mass)


m EG,i
μEG,i = (23.112)
mF

n EG,i MEG,i
μEG,i = (23.113)
n F MF

MEG,i
μEG,i = νEG,i · (23.114)
MF

Elemental Analysis of Fuels

Single-Component Fuel
If the structural chemical formula of a gaseous single-component fuel is known, e.g.
in the form of Ca Hb Oz N p Sr , an elemental analysis to determine the mass fractions
h, c, o, n, s of the fuel can be easily performed. The major advantage is that the
oxygen demand as well as the composition of the exhaust gas can then be calculated
following the simple steps shown in Sect. 23.3.1. The aim is now to investigate the
gaseous fuel Ca Hb Oz N p Sr . Let the molar quantity of the fuel be n F :
• Mass fraction carbon c

The fuel contains a · n F carbon atoms, see structural chemical formula, i.e.

mC n C MC an F MC a MC
c = ξC = = = = (23.115)
mF n F MF n F MCa Hb Oz N p Sr MCa Hb Oz N p Sr

• Mass fraction hydrogen h

The fuel contains b · n F hydrogen atoms, see structural chemical formula, i.e.

bMH
h = ξH = (23.116)
MCa Hb Oz N p Sr
742 23 Combustion Processes

• Mass fraction oxygen n

The fuel contains z · n F oxygen atoms, see structural chemical formula, i.e.

z MO
o = ξO = (23.117)
MCa Hb Oz N p Sr

• Mass fraction nitrogen n

The fuel contains p · n F nitrogen atoms, see structural chemical formula, i.e.

pMN
n = ξN = (23.118)
MCa Hb Oz N p Sr

• Mass fraction sulphur s

The fuel contains r · n F sulphur atoms, see structural chemical formula, i.e.

r MS
s = ξS = (23.119)
MCa Hb Oz N p Sr

Multi-component Fuel (Mixture)


If the fuel is a mixture of several components i, each with a known chemical structural
formula, e.g. methane and propane, an elemental analysis can also be carried out.
The mass of the fuel component i is m i , so that

mi = mF. (23.120)
i

The mass fractions of the carbon of the fuel component i shall be ci , see Fig. 23.5.
The other elements follow accordingly. The elemental analysis for the fuel mixture
is therefore:
• Mass fraction carbon c

ci m i ci m i
i i
c=
= = ci ξi (23.121)
mi mF i
i

• Mass fraction hydrogen h


hi mi
i
h=
= h i ξi (23.122)
mi i
i
23.3 Stoichiometry 743

Fig. 23.5 Exemplary gaseous fuel mixture

• Mass fraction nitrogen n


ni m i
i
n=
= n i ξi (23.123)
mi i
i

• Mass fraction oxygen o


oi m i
i
o=
= oi ξi (23.124)
mi i
i

• Mass fraction sulphur s


si m i
i
s=
= si ξi (23.125)
mi i
i

The mass fractions of the components can be calculated with Eqs. 23.115–23.119.

23.3.5 Setting Up a Chemical Equation

Instead of applying the approach for CHONS-complexes, see Sect. 23.3.2, the chem-
ical equation can be set up alternatively. Note that a chemical equation does not reflect
744 23 Combustion Processes

the mass ratio of the chemical bonds involved,9 but the molar ratio of these bonds.
This approach is illustrated step by step for methanol with λ > 1 as an example10 :
• Stoichiometric with oxygen

3
1 CH3 OH + O2 → 1 CO2 + 2 H2 O (23.126)
2
Left and right hand side of this chemical equation are balanced by one carbon
atom, four hydrogen atoms and four oxygen atoms.
• Stoichiometric with air

3 79 3 79 3
1 CH3 OH + O2 + · N2 → 1 CO2 + 2 H2 O + · N2 (23.127)
2 21 2 21 2
x
The ratio 7921
is the molar ratio xNO2 ,dry air
. It shows how much more nitrogen than
2 ,dry air
oxygen is contained in the combustion air. Since the amount of nitrogen on the left
is equal to the amount of nitrogen on the right, nitrogen is inert in this combustion,
i.e. chemically non-reactive.
• With excess air (dry air)

3 79 3
CH3 OH + λ · O2 + λ · · N2
2 21 2
79 3 3
→ CO2 + 2 H2 O + λ · · N2 + (λ − 1) · O2
21 2 2
(23.128)
In the case λ = 1, there is no oxygen in the exhaust gas, so that the combustion is
then stoichiometric.

According to this chemical Eq. 23.128, both the oxygen demand and the exhaust gas
composition can be determined immediately:
• Oxygen demand

It requires λ · 3
2
times the molar amount of the fuel as amount of oxygen, i.e.

3
n O2 = λ · · nF. (23.129)
2
Thus, it is by division of n F
3
O =λ· (23.130)
2

9 Of course, the mass balance must be satisfied, cf. Sect. 23.3.3.


10 Application of a simplified standard atmosphere containing only nitrogen and oxygen.
23.3 Stoichiometry 745

• Exhaust gas composition

The molar amount of each exhaust component in relation to the molar amount of
fuel can be obtained from Eq. 23.128:
– Carbon dioxide

Comparing the stoichiometric factors of carbon dioxide and methanol in Eq.


23.128 leads to
n EG,CO2 = 1 · n F . (23.131)

Thus, it is by division of n F
νEG,CO2 = 1 (23.132)

This means the molar amount of fuel and carbon dioxide is identical.

– Water

Comparing the stoichiometric factors of water and methanol in Eq. 23.128 leads
to
n EG,H2 O = 2 · n F . (23.133)

Thus, it is by division of n F
νEG,H2 O = 2 (23.134)

This means the molar amount of water is twice as much as the molar amount of
fuel.

– Nitrogen

Comparing the stoichiometric factors of nitrogen and methanol in Eq. 23.128


leads to
79 3
n EG,N2 = λ · · · nF. (23.135)
21 2
Thus, it is by division of n F

79 3
νEG,N2 = λ · · (23.136)
21 2

This means the molar amount of nitrogen is λ · 79


21
· 3
2
times the molar amount
of fuel.

– Oxygen

Comparing the stoichiometric factors of product-oxygen and methanol in Eq.


23.128 leads to
746 23 Combustion Processes

3
n EG,O2 = (λ − 1) · · nF. (23.137)
2
Thus, it is by division of n F

3
νEG,O2 = (λ − 1) · (23.138)
2

This means, in case λ = 1 there is no oxygen in the exhaust gas.

23.3.6 Dew Point of the Exhaust Gas

The analysis of the exhaust gas has shown that fossil fuels emit water during combus-
tion, depending on the amount of hydrogen respectively the amount of water carried
by the fuel and the air. As already known, water tends to change its state of aggre-
gation, i.e. vapour can condense and become liquid. In this section, the conditions
under which water condenses and how much water changes to the liquid state are
investigated. For this purpose, the gaseous exhaust gas is treated as a mixture of ideal
gases. Once the concentration of the exhaust gas is known, the partial pressure of
the vapour in the gas mixture can be calculated, see Chap. 19. Based on the assump-
tion that the entire product water is initially in vapour state, two alternatives can be
pursued:

• Case 1—Molar concentration xEG,i is well-known

This case is simple because the given correlation can be applied to ideal mixtures,
i.e. pi
xi = πi = . (23.139)
p

This equation states that the molar fraction and partial pressure fraction are the
same, see Chap. 19. Hence, the partial pressure of the vapour is

pv = pEG,H2 O = xEG,H2 O · p. (23.140)

• Case 2—Mass concentration ξEG,i is well-known

A conversion from mass-fractions to molar-fractions xEG,i is required, i.e.

MEG
xEG,i = ξEG,i · . (23.141)
MEG,i

Note that MEG is the molar mass of the exhaust gas, i.e. the gaseous mixture. It
can be determined following the principles of Chap. 19. MEG,i , however, is the
23.3 Stoichiometry 747

molar mass of the specific component i with the its individual concentration.11
The partial pressure of the vapour is therefore

MEG
pv = pEG,H2 O = xEG,H2 O · p = ξEG,H2 O · · p. (23.142)
MEG,H2 O

If the partial pressure of the vapour pv in the exhaust gas is known, the associated
saturated steam temperature, i.e. dew point temperature ϑτ , can be looked up in the
steam table, for example, i.e.
ϑτ = f ( pv ) (23.143)

In other words, this is the minimum temperature required to fulfil the assumption that
all water occurs as vapour. If the temperature falls below this threshold, the vapour
begins to condense. In the following, it is analysed how much water condenses
when the temperature drops. This scenario is sketched in Fig. 23.6. As soon as
the condensed water separates from the gas mixture, the remaining exhaust gas is
saturated with vapour according to its temperature, see Chap. 20. Hence, for the
exhaust gas, see Fig. 23.6a, it is

ps (TEG ) ṅ v
πs = = = xs . (23.144)
p ṅ EG − ṅ liq

Obviously, the total molar amount leaving the separator is reduced by the molar
amount of liquid water that has been extracted. The water that is contained in the
remaining exhaust gas is in vapour state. It is known from Chap. 19 that the molar con-
centration is equal to the partial pressure ratio. Setting up the mass balance according
to Fig. 23.6b now results in

ṁ EG,H2 O = ṁ v + ṁ liq . (23.145)

Fig. 23.6 Condensing product water

11 In this section it is the component water.


748 23 Combustion Processes

Dividing by the molar mass of water results in

ṅ EG,H2 O = ṅ v + ṅ liq . (23.146)

A combination of Eqs. 23.144 and 23.146 leads to the molar quantity of the liquid,
i.e. condensed water,
ṅ EG,H2 O − xs ṅ EG
ṅ liq = . (23.147)
1 − xs

The conversion into mass fluxes by multiplication with MH2 O leads to

M H2 O
ṁ EG,H2 O − xs ṁ EG MEG
ṁ liq = . (23.148)
1 − xs

With reference to the mass of the fuel, the following results

M H2 O
ṁ liq μEG,H2 O − xs μEG
= μ∗EG,H2 O =
MEG
(23.149)
ṁ F 1 − xs

where xs is a function of the total pressure p and the saturation pressure ps as a


function of the exhaust gas temperature, see also Eq. 23.144, i.e.

ps (TEG )
xs = (23.150)
p

Problem 23.1 Coal is oxidised with an air fuel equivalence ratio of λ = 1.35. The
air shall be dry. Please calculate the minimum oxygen demand, the overall air demand
and the concentrations ξEG,i of the exhaust gas. The composition of the coal is as
follows: c = 0.75, h = 0.05, s = 0.01, o = 0.06, n = 0.01, w = 0.06, a = 0.06.

Solution

The minimum oxygen demand obeys

omin = 2.667c + 8h + s − o = 2.3502 kgO2 /kgF . (23.151)

The mass fraction of oxygen in a simplified standard atmosphere is ξO2 ,air = 0.23
and the mass fraction of nitrogen is ξN2 ,air = 0.77, so that the minimum air demand
is omin
lmin = = 10.2185. (23.152)
ξO2 ,air

The combustion is performed with excess air, so that the overall air demand follows
23.3 Stoichiometry 749

l = λ · lmin = 13.7949. (23.153)

The exhaust gas composition is


• Carbon dioxide:
μEG,CO2 = 3.667c = 2.7502 (23.154)

• Water:
μEG,H2 O = 9h + w = 0.51 (23.155)

• Sulphur dioxide:
μEG,SO2 = 2s = 0.02 (23.156)

• Nitrogen:
μEG,N2 = n + λlmin ξN2 ,air = 10.6321 (23.157)

• Oxygen:
μEG,O2 = omin (λ − 1) = 0.8226 (23.158)

The mass conservation is satisfied,12 see Sect. 23.3.3:



μEG,i = (1 − a) + λlmin (1 + x) . (23.159)
  
i
   =14.7349
=μEG =14.7349

Hence, the concentration of the exhaust gas is


• Carbon dioxide:
μEG,CO2
ξEG,CO2 = = 0.1866 (23.160)
μEG

• Water:
μEG,H2 O
ξEG,H2 O = = 0.0346 (23.161)
μEG

• Sulphur dioxide:
μEG,SO2
ξEG,SO2 = = 0.0014 (23.162)
μEG

• Nitrogen:
μEG,N2
ξEG,N2 = = 0.7216 (23.163)
μEG

• Oxygen:
μEG,O2
ξEG,O2 = = 0.0558 (23.164)
μEG

12 The air is supposed to be dry so that its water content is x = 0.


750 23 Combustion Processes

Problem 23.2 Pure hydrogen is oxidised with λ = 2.9. Air shall be treated simpli-
fied as a two component mixture of N2 /O2 . The mass-fraction of oxygen shall be
ξO2 ,air = 0.23.
(a) Calculate the minimum air demand lmin and the overall air demand l.
(b) What are the total specific exhaust gas composition μEG and the concentrations
of the exhaust gas components ξEG,i ?
(c) Determine the dew point ϑτ of the exhaust gas at a total pressure of p =
1060 mbar.

Solution

(a) Since the fuel composition is h = 1, the minimum oxygen demand obeys

omin = 2.667c + 8h + s − o = 8. (23.165)

The mass fractions of oxygen respectively nitrogen in a simplified standard


atmosphere are ξO2 ,air = 0.23 and ξN2 ,air = 0.77, so that the minimum air demand
is omin
lmin = = 34.7826. (23.166)
ξO2 ,air

The combustion is performed with excess air, so that the overall air demand
results in
l = λlmin = 100.8696. (23.167)

(b) The exhaust gas composition is


• Water:
μEG,H2 O = 9h + w = 9 (23.168)

• Nitrogen:
μEG,N2 = n + λlmin ξN2 ,air = 77.6696 (23.169)

• Oxygen:
μEG,O2 = omin (λ − 1) = 15.2000 (23.170)

The summation leads to



μEG = μEG,i = 101.8696. (23.171)
i

Hence, the concentrations of the exhaust gas components are


23.3 Stoichiometry 751

• Water:
μEG,H2 O
ξEG,H2 O = = 0.0883 (23.172)
μEG

• Nitrogen:
μEG,N2
ξEG,N2 = = 0.7624 (23.173)
μEG

• Oxygen:
μEG,O2
ξEG,O2 = = 0.1492 (23.174)
μEG

(c) To calculate the dew point ϑτ of the exhaust gas, the partial pressure of the
product water13 is required. The molar fraction xEG,H2 O follows

MEG
xEG,H2 O = ξEG,H2 O . (23.175)
MH2 O

The molar mass of the exhaust gas is

RM
MEG = (23.176)
REG

with
REG = ξEG,i Ri . (23.177)
i

The gas constants of the exhaust gas components are:


• Water
RM J
R H2 O = = 461.9 (23.178)
MH2 O kg K

• Nitrogen
RM J
RN2 = = 296.94 (23.179)
MN2 kg K

• Oxygen
RM J
R O2 = = 259.82 (23.180)
MO2 kg K

Hence, the gas constant of the exhaust gas is


J
REG = ξEG,i Ri = 305.9755 . (23.181)
i
kg K

13 It is assumed that the water is completely vapourised.


752 23 Combustion Processes

Its molar mass is


RM kg
MEG = = 27.1731 . (23.182)
REG kmol

The molar fraction of the vapour in the exhaust gas is

MEG
xEG,H2 O = ξEG,H2 O = 0.1334. (23.183)
MH2 O

As known from Chap. 19 it is


pEG,H2 O
xEG,H2 O = . (23.184)
p

Hence, the partial pressure of the vapour in the exhaust gas results in

pEG,H2 O = xEG,H2 O · p = 0.1414 bar. (23.185)

The saturation temperature corresponding to this pressure can be taken from the
steam table, i.e.
ϑτ = f pEG,H2 O = 52.75 ◦ C. (23.186)

Problem 23.3 Natural gas is oxidised with dry air and an air fuel equivalence ratio
λ = 1.3. What are the minimum oxygen demand Omin , the overall air demand L
and the exhaust gas composition νRG,i ? What are the exhaust gas concentrations in
molar- and mass-fractions? Natural gas shall be treated as a mixture of the following
components:
• Methane CH4 : xCH4 = 0.8
• Ethane C2 H6 : xC2 H6 = 0.02
• Propane C3 H8 : xC3 H8 = 0.01
• Carbon dioxide CO2 : xCO2 = 0.03
• Nitrogen N2 : xN2 = 0.14

Solution

To calculate the minimum oxygen demand, Eq. 23.76 for the example fuel, see Fig.
23.4, is applied and modified, i.e.
bi zi
Omin = 0.5 · xH2 + 0.5 · xCO + (ai + − + ri ) · xCa Hb Oz N p Sr ,i − xO2 .
i
4 2
(23.187)
Methane, ethane and propane are representatives for the CHONS-complexes. Since
the fuel does not contain hydrogen H2 , oxygen O2 and carbon monoxide CO the
23.3 Stoichiometry 753

equation simplifies to
bi zi
Omin = (ai + − + ri ) · xCa Hb Oz N p Sr ,i . (23.188)
i
4 2

Hence, the CHONS-complexes are handled as follows14 :


  
4 6 8
Omin = 1 + · xCH4 + 2 + · xC2 H6 + 3 + · xC3 H8 = 1.72 (23.189)
4 4 4

Thus, 1 kmol of fuel requires 1.72 kmol of oxygen for oxidation. The minimum air
demand is higher because 21 vol.−% of the air consists of oxygen, i.e.
Omin
L min = = 8.1905. (23.190)
xO2 ,air

The overall air demand is

L = λ · L min = 10.6476. (23.191)

The exhaust gas composition can be calculated according to the exemplary fuel, cf.
Fig. 23.4:
• Carbon dioxide CO2

Based on the exemplary mixture from the theory part, i.e.



νEG,CO2 = xCO + ai · xCa Hb Oz N p Sr ,i + xCO2 (23.192)
i

it follows for the given gaseous mixture:

νEG,CO2 = 1 · xCH4 + 2 · xC2 H6 + 3 · xC3 H8 + xCO2 = 0.9 (23.193)

• Water H2 O

The example of the gaseous fuel according to Fig. 23.4 has shown
bi
νEG,H2 O = xH2 + · xCa Hb Oz N p Sr ,i + 1.608 · λ · L min · x. (23.194)
i
2

As the combustion air is dry and the water comes exclusively from the bonded
hydrogen of the fuel, it is

14According to their chemical composition: CH4 → a = 1, b = 4, C2 H6 → a = 2, b = 6 and


C3 H8 → a = 3, b = 8.
754 23 Combustion Processes

νEG,H2 O = 2 · xCH4 + 3 · xC2 H6 + 4 · xC3 H8 = 1.7 (23.195)

• Nitrogen N2

For the nitrogen it is


n EG,N2 pi
νEG,N2 = = x N2 + · xCa Hb Oz N p Sr ,i + λ · L min · xN2 ,dry air . (23.196)
nF i
2

Since the CHONS-complexes do not contain any nitrogen, it simplifies to

νEG,N2 = xN2 + λ · L min · xN2 ,dry air = 8.5516 (23.197)

• Oxygen O2

The example of the gaseous fuel according to Fig. 23.4 has shown
n EG,O2
νEG,O2 = = L min · xO2 ,dry air · (λ − 1) = Omin · (λ − 1). (23.198)
nF

Applied to the composition of the given mixture it is

νEG,O2 = L min · xO2 ,dry air · (λ − 1) = Omin · (λ − 1) = 0.5160 (23.199)

In a next step, the exhaust gas concentration is calculated in molar fractions. To do,
so overall νEG is required, i.e.

νEG = νEG,i = 11.6676. (23.200)
i

Therefore, the concentrations are


• Carbon dioxide:
νEG,CO2
xEG,CO2 = = 0.0771 (23.201)
νEG

• Water:
νEG,H2 O
xEG,H2 O = = 0.1457 (23.202)
νEG

• Nitrogen:
νEG,N2
xEG,N2 = = 0.7329 (23.203)
νEG
23.3 Stoichiometry 755

• Oxygen:
νEG,O2
xEG,O2 = = 0.0442 (23.204)
νEG

To convert these concentrations into mass fractions, the following equation is applied

MEG,i
ξEG,i = xEG,i · . (23.205)
MEG

For this, the molar mass of the exhaust gas is required, i.e.

MEG = xEG,i · Mi
i
= xEG,CO2 MCO2 + xEG,H2 O MH2 O + xEG,N2 MN2 + xEG,O2 MO2 (23.206)
kg
= 27.9541 .
kmol
The mass fractions of the exhaust gas result accordingly
• Carbon dioxide:
MEG,CO2
ξEG,CO2 = xEG,CO2 · = 0.1214 (23.207)
MEG

• Water:
MEG,H2 O
ξEG,H2 O = xEG,H2 O · = 0.0938 (23.208)
MEG

• Nitrogen:
MEG,N2
ξEG,N2 = xEG,N2 · = 0.7341 (23.209)
MEG

• Oxygen:
MEG,O2
ξEG,O2 = xEG,O2 · = 0.0506 (23.210)
MEG

Problem 23.4 A mixture of gaseous fuels contains 75 vol.−% ethane (C2 H6 ) and
25 vol.−% methane (CH4 ). Please conduct an elemental analysis and determine c
and h. What is the minimum oxygen demand omin ?

Solution

In a first step, the elemental analysis of each component, i.e. ethane and methane, is
carried out, i.e.
756 23 Combustion Processes

• Ethane:

Carbon
a MC 2 · MC 24
cC2 H6 = ξC = = = = 0.8 (23.211)
MCa Hb Oz N p Sr MC2 H6 30

and hydrogen

bMH 6 · MH 6
h C2 H6 = ξH = = = = 0.2 (23.212)
MCa Hb Oz N p Sr MC2 H6 30

• Methane:

Carbon
a MC 1 · MC 12
cCH4 = ξC = = = = 0.75 (23.213)
MCa Hb Oz N p Sr MCH4 16

and hydrogen

bMH 4 · MH 4
h CH4 = ξH = = = = 0.25 (23.214)
MCa Hb Oz N p Sr MCH4 16

The molar mass of the gaseous fuel is15


kg
MF = xF,i · Mi = xC2 H6 MC2 H6 + xCH4 MCH4 = 26.5 (23.215)
i
kmol

Now, the composition of the gaseous mixture is converted into mass-fractions, i.e.

MC2 H6
ξC2 H6 = xC2 H6 · = 0.8491 (23.216)
MF

and
MCH4
ξCH4 = xCH4 · = 0.1509. (23.217)
MF

Thus, for the gaseous mixture, the following applies



c= ci ξi = cC2 H6 ξC2 H6 + cCH4 ξCH4 = 0.7925 (23.218)
i

and

15 Note that σi = xi , i.e. molar fractions are identical with volume fractions.
23.3 Stoichiometry 757

Fig. 23.7 Energy


balance—combustion
chamber


h= h i ξi = h C2 H6 ξC2 H6 + h CH4 ξCH4 = 0.2075. (23.219)
i

When the elemental analysis is finished, the minimum oxygen demand can be easily
calculated, i.e.
omin = 2.667c + 8h = 3.7738. (23.220)

23.4 Energetic Balancing

Now that it has been clarified how much air is needed for fossil fuel oxidation and
what the composition of the exhaust gas looks like, the focus in this section is on
the energy balance. Thus, a correlation is derived between the heat released and the
corresponding exhaust gas temperature. However, this correlation is dependent on
the fuel-air mixture. Figure 23.7 shows a combustion chamber as an open system
assumed in steady state. Obviously, fuel and air enter the system while ash and
exhaust gas leave the system, so that the mass balance in steady state obeys

ṁ F + ṁ a = ṁ EG + ṁ Ash . (23.221)

The ash may be disregarded if the fuel is gaseous, so that

ṁ F + ṁ a = ṁ EG . (23.222)

The first law of thermodynamics for steady state states in simplified form that the
energy flux into the system is balanced by the energy flux out of the system, i.e.

   
c2 c2
Q̇ + P + ṁ k · h k + k + g · z k = ṁ i · h i + i + g · z i . (23.223)
in,k
2 out,i
2
758 23 Combustion Processes

Obviously, according to Fig. 23.7, no technical work16 exceeds the system boundary.
Disregarding the ash as well as the kinetic and potential energies, the first law of
thermodynamics follows

Q̇ + ṁ F · h F (TF ) + ṁ a · h a (Ta ) = ṁ EG · h EG (TEG ). (23.224)

By rearranging and substituting the mass of the exhaust gas according to Eq. 23.222
it is
Q̇ = ṁ F [h EG (TEG ) − h F (TF )] + ṁ a [h EG (TEG ) − h a (Ta )] . (23.225)

The specific enthalpy is only a function of temperature, since fuel, air and exhaust
gas are treated as ideal fluids17 in a simplified way. TF is the temperature of the fuel,
Ta the temperature of the combustion air and TEG the exhaust gas temperature. Open
systems have been balanced several times in the previous Parts and it is expected that
one can replace the specific enthalpies by applying a caloric equation of state, e.g.

h 2 − h 1 = c p (T2 − T1 ) . (23.226)

Unfortunately, an application of the caloric equation of state in Eq. 23.225 fails,


because in the first bracket the released exhaust gas and the supplied fuel are com-
pared, respectively exhaust gas and supplied air in the second bracket.18 Until now,
enthalpy has included the caloric effect, i.e. the temperature change depending on the
supplied heat,19 but chemical binding energies are not part of the specific enthalpy.
Consequently, the approaches that have been used so far are limited as the energy
content of different elements and compounds cannot be compared with each other.
However, this dilemma is solved in Chap. 24, where the absolute enthalpies are
introduced, which include not only the caloric part but also the chemically bonded
energy. In this chapter, a different pragmatic approach is taken, which divides the
combustion process into three conceptual, sequential steps.

23.4.1 Lower Heating Value

The major problem with the specific enthalpy applied so far is that the conversion of
substances is not covered energetically. The systems dealt with in Parts I/II have been
ideal/real fluids, which, however, are not chemically active. Consequently, the chem-
ical reaction and thus the conversion of substances must be described alternatively,

16 Mechanical or electrical energy.


17 The pressure dependence of the specific enthalpy for incompressible liquids is neglected.
18 The enthalpy differences cannot be replaced by temperature differences because exhaust gas and

fuel/air have different specific heat capacities.


19 The supplied thermal energy increases the temperature and thus the enthalpy, while the heat

release decreases the temperature and the enthalpy.


23.4 Energetic Balancing 759

Fig. 23.8 Definition of the


lower heating value HU

see Fig. 23.8. Both fuel and air enter the combustion chamber with the same refer-
ence temperature T0 . During oxidation, the temperature rises so that under adiabatic
conditions the exhaust gas exits at a higher temperature than the inlet temperature.
Let us now drop this adiabatic condition and apply cooling. The cooling should be
so effective that the exhaust gas exits with the same reference temperature T0 as the
reactants have been supplied. The corresponding cooling power shall be

Q̇ = ṁ F q < 0. (23.227)

Obviously, the released heat flux depends on the fluid, so that the so-called lower
heating value HU (T0 ) > 0 is defined as follows

!
Q̇ = ṁ F q = −ṁ F HU (T0 ) < 0. (23.228)

Definition 23.1 The lower heating value HU (T0 ) of fuels is the specific amount
of heat required to cool the exhaust gas back to a reference temperature T0 of the
supplied fuel and air. Combustion is therefore isothermal at T0 . By definition, the
entire product water is in a vapour state in this case.

In order to determine the lower heating value experimentally, it must be ensured that
oxidation is complete, i.e. that the highest oxidation level is reached. Therefore, the
experiments are carried out with an air-fuel equivalence ratio of λ > 1. The excess
air, which is chemically non-reactive, has no energetic influence on the heating value,
since the air inlet and outlet temperatures are identical T0 and thus have no caloric
effect. Under the same premise, humid, unsaturated air can be supplied because the
moisture it contains, i.e. the vapour, has the same inlet and outlet temperature.20
Tables 23.2 and 23.3 show lower heating values of relevant technical fossil fuels.

20 Therefore, Fig. 23.8 also works with unsaturated humid air ṁ A . If liquid water were supplied
with the combustion air, thermal energy would be required to vaporise the liquid water and turn it
into vapour. This would reduce the specific lower heating value.
760 23 Combustion Processes

Table 23.2 Lower heating value HU (solids and liquids) at ϑ0 = 25 ◦ C, see [10]
Solid fuels Liquid fuels
Fuel HU in MJ
kg Fuel HU in MJ
kg
Wood, dry 14.65 . . . 16.75 Ethanol 26.9
Turf, dry 11.72 . . . 15.07 Benzol 40.15
Raw lignite 8.37 . . . 11.30 Toluol 40.82
Brown coal briquettes 19.68 . . . 20.10 Naphthalene 38.94
Hard coal 27.31 . . . 34.12 Pentane 45.43
Anthracite 32.66 . . . 33.91 Octane 44.59
Coke 27.84 . . . 30.35 Benzine 42.7

Table 23.3 Lower heating value HU (Gases) at ϑ0 = 25 ◦ C, see [10]


Gaseous fuels
Fuel HU in MJ
kg
Hydrogen 119.97
Carbon monoxide 10.10
Methane 50.01
Ethane 47.49
Propane 46.35
Ethylene 47.15
Acetylene 48.22

23.4.2 Conceptual 3-Steps Combustion

As discussed, the chemically reactive part of the combustion is treated energetically


with the lower heating value HU (T0 ). However, the inlet and outlet temperatures
usually differ from the reference temperature T0 to which the lower heating value
is referred, see Fig. 23.7. To calculate a non-isothermal combustion process with
different inlet and outlet temperatures, pre- and post-conditioning is required in
addition to the lower heating value. This concept is explained in more detail in this
section.

Step 1—Pre-conditioning

Fuel and air enter the combustion chamber at any temperature TF = T0 and Ta = T0
respectively. In the chamber, both fluids are conditioned to the reference temperature
T0 so that the lower heating value approach can be followed in step 2. In step 1, cf.
Fig. 23.9, no chemical reactions take place and the energy balance is based purely
on caloric effects. Consequently, the first law of thermodynamics obeys
23.4 Energetic Balancing 761

Fig. 23.9 Combustion—


Step 1

Q̇ 1 + ṁ F h F (TF ) + ṁ a h a (Ta ) = ṁ F h F (T0 ) + ṁ a h a (T0 ) . (23.229)

Solving for the required heat flux results in

Q̇ 1 = −ṁ F [h F (TF ) − h F (T0 )] − ṁ a [h a (Ta ) − h a (T0 )] . (23.230)

In specific notation with reference to the mass flux of the fuel it is

q1 = − [h F (TF ) − h F (T0 )] − λlmin [h a (Ta ) − h a (T0 )] (23.231)

with
ṁ a
λlmin = l = . (23.232)
ṁ F

This equation can actually be calculated with the means of Part I: the caloric equation
of state can be applied.21 The specific heat flux q1 can be positive, negative or zero
depending on the temperatures TF and Ta compared to the reference temperature T0 .

Step 2—Chemical Reaction

Since the fluids are now both conditioned to the reference temperature T0 , the chemi-
cal reaction can be implemented energetically using the lower heating value approach,
see Fig. 23.10. The exhaust gas also leaves step 2 at reference temperature. Thus, the
cooling power follows, cf. Sect. 23.4.1,

Q̇ 2 = −ṁ F HU (T0 ) . (23.233)

In specific notation with reference to the mass flux of the fuel, the following results

q2 = −HU (T0 ) (23.234)

21 The first bracket in Eq. 23.231 contains pure fuel, the second pure air.
762 23 Combustion Processes

Fig. 23.10 Combustion—


Step 2

Fig. 23.11 Combustion—


Step 3

Note that the exhaust gas leaves step 2 with reference temperature T0 . Due to the
definition of the lower heating value, cf. Definition 23.1, the product water is in the
gaseous state when leaving step 2.

Step 3—Post-conditioning

The exhaust gas leaves the combustion chamber with a temperature TEG that is usually
not equal to the reference temperature T0 , so that post-conditioning must take place
in the equivalent model, see Fig. 23.11. Since there is no chemical reaction, this is
merely a caloric step. In this case, the first law of thermodynamics in steady state
yields

Q̇ 3 + ṁ EG h EG (T0 ) + ṁ Ash h Ash (T0 ) = ṁ EG h EG (TEG ) + ṁ Ash h Ash (TEG ) .


(23.235)
Solving for the heat flux

Q̇ 3 = ṁ EG [h EG (TEG ) − h EG (T0 )] + ṁ Ash [h Ash (TEG ) − h Ash (T0 )] . (23.236)

In specific notation with reference to the mass flux of the fuel, the following results
23.4 Energetic Balancing 763


q3 = μEG,i · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] (23.237)
i

with
ṁ EG
μEG,i = μEG = (23.238)
i
ṁ F

and
ṁ Ash
a= . (23.239)
ṁ F

This energy balance can actually be calculated with the means of Part I: The caloric
equation of state can be applied if the specific heat capacity of the exhaust gas is
known.

Conclusion

Steps 1–3 can be summarised so that the entire combustion process is covered by
the following energy balance, which is a combination of Eqs. 23.231, 23.234 and
23.237:

q = q1 + q2 + q3 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] + (23.240)
− [h F (TF ) − h F (T0 )] − λlmin · [h a (Ta ) − h a (T0 )] − HU (T0 ).

A superposition of the combustion process can be visualised according to Fig. 23.12.


Thermal energy Q = Q 1 + Q 2 + Q 3 is transferred, which includes both caloric and
chemical effects. The overall system boundary leads to the following mass fluxes:
• Inlet: Fuel @ TF and dry air @ Ta
• Outlet: Exhaust gas @ TEG , Ashes22 @ TEG
Although the combustion air has been assumed to be dry, water may be present in
the exhaust gas, e.g. water bound in the fuel or hydrogen that is oxidised, i.e. due to
w and h. The exhaust gas shown in Fig. 23.12 may therefore contain water, but this
water is in the vapour state, as this is the premise of step 2: applying the lower heating
value approach means that the product is water in the vapour state. This vapour has
been treated in step 3 in the form of sensible heat only, no latent heat associated
with a phase change has been applied. However, the exhaust gas temperature may be
below the dew point and liquid water occurs, see Sect. 23.3.6. The energy balance
can then be supplemented by a step 4, see Fig. 23.13. Both the ash and the exhaust
gas23 leave step 3. Let us now split the exhaust gas leaving step 3 into two parts: one

22 Ashes normally occur with solid fuels. The effect of ashes on the energy balance disappears for

a = 0.
23 With its water in vapour state.
764 23 Combustion Processes

Fig. 23.12 Combustion—summary

Fig. 23.13 Condensation of water

containing all the water in vapour state ṁ EG,H2 O , the other containing all the other
components except the water, i.e.

ṁ EG,dry = ṁ EG − ṁ EG,H2 O . (23.241)

The focus is now on the water: the mass of liquid water ṁ liq when the exhaust gas
temperature is below the dew point can be determined according to Sect. 23.3.6. The
rest of the water is still in vapour state, so that the remaining exhaust gas has a mass of

ṁ EG = ṁ EG − ṁ liq . (23.242)

This mass ṁ EG contains the remaining vapour and dry exhaust gas. Let us now apply
the first law of thermodynamics in the steady state, which follows the system bound-
ary of Fig. 23.13, which includes the phase change due to condensation. Water in
vapour state as well as a heat flux Q̇ 4 enter the system, while liquid water and water
in vapour state leave the system, i.e. with the other energy fluxes it yields

ṁ EG,H2 O h v (TEG ) + Q̇ 4 + ṁ EG,dry h EG (TEG ) + ṁ Ash h Ash (TEG ) =



ṁ liq h liq (TEG , p) + ṁ EG,H2 O − ṁ liq h v (TEG ) + (23.243)
+ ṁ EG,dry h EG (TEG ) + ṁ Ash h Ash (TEG ) .
23.4 Energetic Balancing 765

The specific enthalpies of vapour and liquid can be taken from the steam table, cf.
Appendix A. For the heat of condensation Q̇ 4 the following therefore applies24 :
 
Q̇ 4 = ṁ liq h liq (TEG , p) − h v (TEG ) ≈ −ṁ liq h v (TEG ) . (23.244)

Since step 4 is isothermal, only the phase change of the water is energetically relevant.
In specific notation, the first law of thermodynamics applies to this step as follows

ṁ liq
q4 = − h v (TEG ) (23.245)
ṁ F

The total specific heat when condensation occurs is therefore

q = q1 + q2 + q3 + q4 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
− [h F (TF ) − h F (T0 )] − λlmin · [h a (Ta ) − h a (T0 )] − HU (T0 )+ (23.246)
ṁ liq
− h v (TEG )
ṁ F

The mass of the liquid ṁ liq follows according to Sect. 23.3.6.


Caloric Equations of State
To solve Eq. 23.240, the specific enthalpies must be calculated. With the assumption
that all components are ideal, the specific enthalpies result:
• Dry Air

h a (Ta ) − h a (T0 ) = c p,a |ϑϑa0 · (ϑa − ϑ0 ) = c p,a |ϑ0 a · ϑa − c p,a |ϑ0 0 · ϑ0 (23.247)

• Exhaust gas

h EG (TEG ) − h EG (T0 ) = c p,EG |ϑ0 EG · ϑEG − c p,EG |ϑ0 0 · ϑ0 (23.248)

The exhaust gas is to be regarded as a mixture of ideal gases whose concentrations


are known. In Chap. 19 it has been shown that the specific heat capacity of such a
mixture is as follows
c p,total = ξi c p,i . (23.249)
i

24The enthalpy difference between the supercooled liquid and the saturated liquid is assumed to
be negligible. A consideration of this aspect is given in Sect. 23.4.3.
766 23 Combustion Processes

Table 23.4 Averaged specific heat capacity c p |ϑ0 in kJ


kg K , according to [5]
ϑ [◦ C] Air N2 O2 CO2 H2 O SO2
–25 1.0034 1.0393 0.9135 0.8035 1.8567 0.6010
0 1.0037 1.0394 0.9147 0.8173 1.8589 0.6079
25 1.0042 1.0395 0.9163 0.8307 1.8615 0.6149
50 1.0048 1.0397 0.9182 0.8437 1.8646 0.6219
75 1.0055 1.0400 0.9204 0.8563 1.8682 0.6289
100 1.0064 1.0404 0.9229 0.8684 1.8724 0.6359
150 1.0087 1.0416 0.9288 0.8914 1.8820 0.6495
200 1.0116 1.0435 0.9354 0.9128 1.8931 0.6626
400 1.0285 1.0567 0.9649 0.9856 1.9466 0.7078
600 1.0498 1.0763 0.9925 1.0427 2.0083 0.7418
800 1.0712 1.0976 1.0158 1.0885 2.0744 0.7672
1000 1.0910 1.1179 1.0350 1.1257 2.1421 0.7867
2000 1.1614 1.1914 1.0991 1.2378 2.4425 0.8410

Hence, for the exhaust gas it is



c p,EG |ϑ0 = ξEG,i · c p,i |ϑ0 . (23.250)
i

Where ξEG,i is the mass fraction of a component i of the exhaust gas and c p,i |ϑ0 is
the temperature-averaged specific heat capacity of this component.
• Fuel
h F (TF ) − h F (T0 ) = c p,F |ϑ0 F · ϑF − c p,F |ϑ0 0 · ϑ0 (23.251)

• Ashes

h Ash (TEG ) − h Ash (T0 ) = cAsh |ϑ0 EG · ϑEG − cAsh |ϑ0 0 · ϑ0 (23.252)

The specific heat capacities are usually temperature-dependent, see Sect. 12.4.4.
Table 23.4 gives an overview of the most relevant components.

Conversion HU (T0 ) → HU (T )

The specific lower heating value is temperature dependent, so a reference temperature


T0 is given at which its value has been determined. In Table 23.2 it is a reference
temperature of ϑ0 = 25 ◦ C that is commonly chosen. Occasionally, however, it is
necessary to convert a given specific lower heating value from a reference temperature
T0 to another temperature T , cf. Fig. 23.14. By definition of the lower heating value,
inlet and outlet temperature are equal, i.e. in this case T . Thus, the converted specific
lower heating value results according to Fig. 23.14, i.e.
23.4 Energetic Balancing 767

Fig. 23.14 HU (T0 ) → HU (T )

!
− ṁ F HU (T ) = Q̇ 1 + Q̇ 2 + Q̇ 3 . (23.253)

Applying Eq. 23.240 leads to

−ṁ F HU (T ) = − ṁ F [h F (T ) − h F (T0 )] − ṁ a [h a (T ) − h a (T0 )] +


− ṁ F HU (T0 ) + ṁ F μEG · [h EG (T ) − h EG (T0 )] + (23.254)
+ ṁ F a · [h Ash (T ) − h Ash (T0 )] .

The solution for HU (T ) yields the specific lower heating value at the new temperature
T , i.e.

HU (T ) = HU (T0 ) + [h F (T ) − h F (T0 )] + λlmin [h a (T ) − h a (T0 )] +


(23.255)
− μEG [h EG (T ) − h EG (T0 )] − a [h Ash (T ) − h Ash (T0 )]

Combustion with Humid Air

In this section, a distinction is made between unsaturated moist combustion air and
combustion air containing water in a non-vapour state.
• Humid, unsaturated air:

This case is shown in Fig. 23.15, assuming that the humid air is unsaturated before
and after preconditioning in step 1, i.e. all the water is in a vapour state throughout
the combustion process. The first law of thermodynamics, including steps 1–3, is
then according to Eq. 23.240
768 23 Combustion Processes

Fig. 23.15 Combustion with unsaturated humid air—summary

Q̇ = Q̇ 1 + Q̇ 2 + Q̇ 3 =
ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] +
 
− ṁ F [h F (TF ) − h F (T0 )] − ṁ a · h 1+x (Ta , x) − h 1+x (T0 , x) +
− ṁ F HU (T0 ).
(23.256)
Note that the specific enthalpy of the humid air h 1+x follows Chap. 20, i.e. it
is referred to the mass of the dry air ṁ a and not to the mass of the humid air
ṁ A = (1 + x) ṁ a . In specific notation with reference to the mass flux of the fuel,
the first law of thermodynamics then obeys

q = q1 + q2 + q3 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
  (23.257)
− [h F (TF ) − h F (T0 )] − λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) +
− HU (T0 ).

If the temperature after step 3 is below the dew point temperature, another step 4
according to Fig. 23.13 must be added to include condensation in the calculation,
i.e.
Q̇ 4 = −ṁ liq h v (TEG ) < 0. (23.258)

The condensed water can be calculated according to Sect. 23.3.6. Thus, the entire
energy balance follows

Q̇ = Q̇ 1 + Q̇ 2 + Q̇ 3 + Q̇ 4 =
ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] +
 
− ṁ F [h F (TF ) − h F (T0 )] − ṁ a · h 1+x (Ta , x) − h 1+x (T0 , x) +
− ṁ F HU (T0 ) − ṁ liq h v (TEG ) .
(23.259)
23.4 Energetic Balancing 769

Fig. 23.16 Combustion with saturated humid air—summary

In specific notation with reference to the mass flux of the fuel, the first law of
thermodynamics then obeys

q = q1 + q2 + q3 + q4 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
 
− [h F (TF ) − h F (T0 )] − λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) + (23.260)
ṁ liq
− HU (T0 ) − h v (TEG ) .
ṁ F

Note that the enthalpy difference [h EG (TEG ) − h EG (T0 )] is related to step 3, so that
the exhaust gas composition of step 3 is relevant.

• Humid, saturated air with liquid water:

This case is shown in Fig. 23.16. Both the fuel and the humid air enter step 1.
Humid air contains dry air (a), vapour (v) as well as liquid water (liq). During
preconditioning in step 1, the total mass of water remains constant, but the mass
of vapour may decrease while the mass of liquid water increases.25 The latent heat
due to condensation in step 1 is part of the specific enthalpies of the humid air h 1+x
and does not need to be treated separately. As usual, the chemical reaction is treated
in step 2: The lower heating value approach only includes the heat released during
oxidation with the product water in vapour state, i.e. it is based on the reaction
of fuel (F) and dry air (a). The available vapour (v) is energetically inert, as it
enters and leaves at the same temperature. The liquid water (liq) is ignored in the
approach to the lower heating value, so it enters and leaves step 2. Excess air, i.e.

25E.g. lowering the temperature from Ta to T0 decreases the capability of air to absorb vapour, see
Chap. 20. In case Ta < T0 the amount of liquid water sinks and the amount of vapour rises.
770 23 Combustion Processes

λ > 1, and vapour are now counted as part of the exhaust gas (EG) and enter step 3.
In step 3, the vaporisation of the liquid water that has been supplied to the system
with the air takes place: The mass of the liquid vaporised in stage 3 is therefore

ṁ liq,a = ṁ a xliq (23.261)

with xliq being the liquid water content of the air after step 1. The vaporisation of
that liquid comes along with a supply of heat Q̇ 3 > 0, i.e.
 
Q̇ 3 = ṁ liq,a h v (T0 ) − h liq (T0 , p) ≈ ṁ liq,a h v (T0 ) . (23.262)

In step 4, the post-conditioning is carried out: Ash and exhaust gas are post-
conditioned to the exhaust gas temperature as in the previous examples. The
exemplary first law of thermodynamics then reads as follows

Q̇ = Q̇ 1 + Q̇ 2 + Q̇ 3 + Q̇ 4 =
ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] +
 
− ṁ F [h F (TF ) − h F (T0 )] − ṁ a h 1+x (Ta , x) − h 1+x (T0 , x) +
− ṁ F HU (T0 ) + ṁ a xliq (T0 ) h v (T0 ) .
(23.263)
In specific notation with reference to the mass flux of the fuel the first law of
thermodynamics yields

q =q1 + q2 + q3 + q4 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
  (23.264)
− [h F (TF ) − h F (T0 )] − λlmin h 1+x (Ta , x) − h 1+x (T0 , x) +
− HU (T0 ) + λlmin xliq (T0 ) h v (T0 ) .

If the temperature after step 4 is below the dew point temperature, another step 5
according to Fig. 23.13 must be added to include condensation in the calculation,
i.e.  
Q̇ 5 = ṁ liq h liq (TEG , p) − h v (TEG ) ≈ −ṁ liq h v (TEG ) . (23.265)

Hence, the entire energy balance in specific notation is

q = q1 + q2 + q3 + q4 + q5 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
 
− [h F (TF ) − h F (T0 )] − λlmin h 1+x (Ta , x) − h 1+x (T0 , x) + (23.266)
ṁ liq
− HU (T0 ) + λlmin xliq (T0 ) h v (T0 ) − h v (TEG )
ṁ F
23.4 Energetic Balancing 771

Fig. 23.17 Definition of the higher heating value H0

Note that the enthalpy difference [h EG (TEG ) − h EG (T0 )] is related to step 4, so the
exhaust gas composition of step 4 is relevant. The mass of the condensing water
follows, see Sect. 23.3.6,
M H2 O
ṁ liq μEG,H2 O − xs μEG
= μ∗EG,H2 O =
MEG
(23.267)
ṁ F 1 − xs

where xs is a function of the total pressure p and the saturation pressure ps as a


function of the exhaust gas temperature, see also Eq. 23.144, i.e.

ps (TEG )
xs = . (23.268)
p

23.4.3 Higher Heating Value

The lower heating value is the specific heat release of an oxidising fuel at constant
reactant and product temperature T0 under the condition that the possible product
water26 is gaseous, i.e. no water condenses, see Sect. 23.4.1. As the previous sections
have shown, the thermal energy released from a combustion increases when gaseous
water condenses and changes to the liquid state. Based on this phase change energy,
the specific higher heating value is defined, cf. Fig. 23.17. In contrast to the specific
lower heating value, the specific upper heating value H0 (T0 ) is the specific amount
of heat released isothermally at T0 when the produced water is completely in liquid
state.
Definition 23.2 The upper heating value H0 (T0 ) of fuels is the specific amount
of heat required to cool the exhaust gas back to a reference temperature T0 of the

26 According to Fig. 23.17, the water originates from the fuel: it can contain water or water can be

a result of the oxidation of hydrogen.


772 23 Combustion Processes

supplied fuel and air. Combustion is therefore isothermal at T0 . By definition, the


entire product water is in liquid state in this case.
After step 1, in which the chemical reaction takes place at T0 and the product water
is gaseous, i.e. when using the lower heating value approach, step 2 follows, which
involves condensation. The heat released in step 1 is already known and follows

Q̇ 1 = −ṁ F HU (T0 ) < 0. (23.269)

In step 2, the exhaust gas is split into dry exhaust gas ṁ EG,dry , which contains no
water, and the remaining water. Water in vapour state ṁ EG,H2 O enters step 2 and
liquid water ṁ liq is present at the exit. The mass balance for the water is

ṁ EG,H2 O = ṁ liq . (23.270)

The first law of thermodynamics in steady state for step 2 obeys

Q̇ 2 + ṁ EG,H2 O h v (T0 ) + ṁ EG,dry h EG,dry (T0 ) + ṁ Ash h Ash (T0 ) =


ṁ EG,dry h EG,dry (T0 ) + ṁ Ash h Ash (T0 ) + ṁ liq h liq (T0 , p).
(23.271)
Solving for Q̇ 2 results in
 
Q̇ 2 = ṁ EG,H2 O h liq (T0 , p) − h v (T0 ) . (23.272)

The specific enthalpies of vapour and liquid can be taken from the steam table, cf.
Sect. A. With a reference temperature for the heating value of ϑ0 = 25 ◦ C for instance
and a pressure27 of p = 1 bar it is
• Vapour (treated as an ideal gas)

kJ
h v (ϑ0 ) = h  (ϑ0 ) = 2546.5 (23.273)
kg

Or alternatively, following Chap. 20

kJ
h v (ϑ0 ) = h v (0 ◦ C) + c p,v · ϑ0 = 2546.5 (23.274)
kg

• Supercooled liquid
kJ
h liq (ϑ0 , p) = 104.9281 (23.275)
kg

Thus, it is

27Note that the pressure of the liquid is equal to the pressure of the gaseous atmosphere above the
water.
23.4 Energetic Balancing 773

Table 23.5 Higher heating values H0 at ϑ0 = 25 ◦ C, according to [10]


Solid fuels Liquid fuels
Fuel H0 in MJ
kg Fuel H0 in MJ
kg
Wood, dry 15.91 . . . 18.0 Ethanol 29.73
Turf, dry 13.83 . . . 16.33 Benzol 41.87
Raw lignite 10.47 . . . 12.98 Toluol 42.75
Brown coal briquettes 20.93 . . . 21.35 Naphthalene 40.36
Hard coal 29.31 . . . 35.17 Pentane 49.19
Anthracite 33.49 . . . 34.75 Octane 48.15
Coke 28.05 . . . 30.56 Benzine 46.05

  kJ kJ
h v (T0 ) − h liq (T0 , p) = 2441.6 ≈ h v (ϑ0 ) = 2441.7 . (23.276)
kg kg

With this simplification28 the heat release in step 2 is


 
Q̇ 2 = ṁ EG,H2 O h liq (T0 , p) − h v (T0 ) ≈ −ṁ EG,H2 O h v (ϑ0 ) < 0. (23.277)

By definition, the higher heating value is

!
Q̇ 1 + Q̇ 2 = −ṁ F H0 (T0 ). (23.278)

Solving for H0 (T0 ) leads to

ṁ EG,H2 O
H0 (T0 ) = HU (T0 ) + h v (T0 ). (23.279)
ṁ F
ṁ EG,H2 O
With μEG,H2 O = ṁ F
one finally gets

H0 (T0 ) = HU (T0 ) + μEG,H2 O · h v (T0 ) (23.280)

Note that the product water is due to w and h, i.e. it comes purely from the fuel and
not from the moisture in the combustion air, see [11]. If humid air is used, the water
content in the exhaust gas increases and its condensation leads to an increase in the
specific heating value. For this reason, Fig. 23.17 emphasises that dry air ṁ a is used.
Specific higher heating values for technically relevant fuels are given in Tables 23.5
and 23.6.

28The enthalpy difference between saturated liquid and supercooled liquid is assumed to be negli-
gible. Note that the vapour is treated as an ideal gas, i.e. the enthalpy is a function of temperature
exclusively. The pressure dependence of the enthalpy of an ideal incompressible liquid has been
neglected.
774 23 Combustion Processes

Table 23.6 Higher heating values H0 (Gases) at ϑ0 = 25 ◦ C, according to [10]


Gaseous fuels
Fuel H0 in MJ
kg
Hydrogen 141.80
Carbon monoxide 10.10
Methane 55.50
Ethane 51.88
Propane 50.35
Ethylene 50.28
Acetylene 49.91

Fig. 23.18 Combustion with saturated humid air—applying the specific higher heating value

Example 23.1 The most generic case of a combustion with saturated humid air has
been introduced based on the specific lower heating value HU (T0 ), see Fig. 23.16 and
the corresponding Eq. 23.266. However, it can also be carried out with the specific
higher heating value H0 (T0 ), cf. Fig. 23.18. In doing so, step 2 differs from the
approach with the lower heating value: The combustion based on the specific higher
heating value requires a combustion with dry air (a) and the product water is entirely
in liquid state, i.e.
ṁ liq,2 = ṁ EG,H2 O . (23.281)

Energetically, the heat released in step 2 under these conditions is

Q̇ 2 = −ṁ F H0 (T0 ) . (23.282)

In step 3, the entire liquid water, i.e. (liq, 2) and (liq, a), is vaporised by heat supply
Q̇ 3 , i.e.

Q̇ 3 = ṁ liq,2 + ṁ liq,a h v (T0 ) = ṁ EG,H2 O + ṁ liq,a h v (T0 ) (23.283)
23.4 Energetic Balancing 775

with, cf. Eq. 23.261,


ṁ liq,a = ṁ a xliq (T0 ) . (23.284)

The following steps are the same as those shown in Fig. 23.16. In step 4, the exhaust
gas is lifted to the exhaust gas temperature TEG with the water being in gaseous state.
If the exhaust gas temperature is below the dew point, step 5 is needed to capture the
condensation. The total energy balance is therefore the result of steps 1–5 and is as
follows

q = q1 + q2 + q3 + q4 + q5 =
μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +
 
− [h F (TF ) − h F (T0 )] − λlmin h 1+x (Ta , x) − h 1+x (T0 , x) +
−H0 (T0 ) + μEG,H2 O h v (T0 ) +λlmin xliq (T0 ) h v (T0 ) + (23.285)
  
−HU (T0 )
ṁ liq
− h v (TEG ) .
ṁ F

Hence, Eqs. 23.266 and 23.285 lead to the same result.

23.4.4 Combustion Calculation Component by Component

Instead of following the humid air approach as before, the individual reactants can
also be calculated step by step. This procedure is illustrated in Fig. 23.19 and is
particularly beneficial when the combustion air has an arbitrary composition. In step
1, the reactants are pre-conditioned to reference temperature T0 . The first law of
thermodynamics for step 1 yields


n
Q̇ 1 = ṁ F [h F (T0 ) − h F (TF )] + ṁ a,i [h i (T0 ) − h i (Ta )] (23.286)
i=1

respectively with29 ṁ a,i = ṁ a ξa,i


n
Q̇ 1 = ṁ F [h F (T0 ) − h F (TF )] + ṁ a ξa,i [h i (T0 ) − h i (Ta )] . (23.287)
i=1

This step only considers the sensible heat,30 so that the caloric equations of state for
step 1 results in
h F (T0 ) − h F (TF ) = c p,F |ϑϑF0 (ϑ0 − ϑF ) (23.288)

29 The composition of air is given by ξa,i .


30 I.e. each reactant is brought to T0 without any phase change being treated.
776 23 Combustion Processes

Fig. 23.19 Component-wise calculation

and
h i (T0 ) − h i (Ta ) = c p,i |ϑϑa0 (ϑ0 − ϑa ) . (23.289)

It should be noted that, in case the vapour of the air condenses during precondition-
ing, additionally latent heat ṁ ∗liq h v (T0 ) is released. This phase change is treated
by step 2
Q̇ 2 = −ṁ ∗liq h v (T0 ) (23.290)

where ṁ ∗liq is the water mass flux that condenses. In step 3, the chemical reaction takes
place. When applying the lower heating value approach the water in the exhaust gas
is entirely in vapour state. Furthermore, the water in the air after step 2 (vapour and
liquid state) is energetically not active, since the temperature remains T0 = const. in
step 3. Hence, the energy balance yields

Q̇ 3 = −ṁ F HU (T0 ) . (23.291)

In case TEG is above dew point temperature, step 4 is required to vaporise the remain-
ing liquid water. However, this is done isothermally at T0 , so that the supplied latent
heat is
Q̇ 4 = ṁ ∗liq h v (T0 ) = − Q̇ 2 . (23.292)

Finally, in step 5, the exhaust gas temperature is achieved by

Q̇ 5 = ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] . (23.293)

This obviously results in the total heat exchanged:


23.4 Energetic Balancing 777

Q̇ = Q̇ 1 + Q̇ 2 + Q̇ 2 + Q̇ 4 + Q̇ 5 =
= ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] +

n
− ṁ F HU (T0 ) + ṁ F [h F (T0 ) − h F (TF )] + ṁ a ξa,i [h i (T0 ) − h i (Ta )]
i=1
(23.294)
Hence, the latent heat ṁ ∗liq h v (T0 ) disappears, as it is released in step 2 and supplied
in step 4. In specific notation the balance results in

q = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +



n
(23.295)
− HU (T0 ) + [h F (T0 ) − h F (TF )] + l · ξa,i [h i (T0 ) − h i (Ta )] .
i=1

Applying the caloric equation of state, the following equation is obtained for ideal
fluids31

q = μEG · c p,EG |ϑϑEG


0
(ϑEG − ϑ0 ) + a · cAsh |ϑϑEG
0
(ϑEG − ϑ0 ) +

n
(23.296)
− HU (T0 ) − c p,F |ϑϑF0 (ϑF − ϑ0 ) − l · ξa,i · c p,i |ϑϑa0 (ϑa − ϑ0 )
i=1

According to Fig. 23.19, Eq. 23.296 can be applied as long as the exhaust gas tem-
perature at the outlet of step 5 is greater than the dew point temperature. This also
applies if water condenses in step 1/2, i.e. while pre-conditioning. However, if the
exhaust gas temperature after step 5 is below the dew point temperature, the heat of
vaporisation of the condensed water must be taken into account in a further step 6.

23.4.5 Molar and Volume Specific Lower/Higher Heating


Value

So far, the heating values HU (T0 ) respectively H0 (T0 ) have been given in relation
to the mass of the fuel, i.e. in [ kg
kJ
]. Additionally, it makes sense, to give the specific
heating values with reference to the molar quantity, i.e. in [ kmolkJ
], respectively with
reference to the volume, i.e. in [ m3 ]:
kJ

• Molar specific lower/higher heating value in [ kmol


kJ
]

The overall heating value is

31 Both the exhaust gas and gaseous fuels are treated as ideal gas. If the fuel is solid or liquid,
respectively, it is assumed to be incompressible, while the pressure dependence of the specific
enthalpy is neglected. The same applies to the ash.
778 23 Combustion Processes

m F · HU (T0 ) = n F HUM (T0 ) , (23.297)

so that its conversion from the mass specific lower heating value HU (T0 ) to the
molar specific lower heating value HUM (T0 ) follows

HUM (T0 ) = MF · HU (T0 ) (23.298)

The molar specific higher heating value H0M (T0 ) follows accordingly

H0M (T0 ) = MF · H0 (T0 ) (23.299)

In ideal fuel mixtures, the molar lower/higher heating values apply according to
the rule of mixtures, i.e.
ni
HUM (T0 ) = · HUM,i (T0 ) (23.300)
i
nF

respectively
ni
H0M (T0 ) = · H0M,i (T0 ) (23.301)
i
nF

• Lower/higher heating value with reference to the volume in [ mkJ3 ]

The overall heating value is

n F · HUM (T0 ) = VF HUv (T0 ) , (23.302)

so that its conversion from the molar specific lower heating value HUM (T0 ) to the
volume specific lower heating value HUv (T0 ) follows for standard conditions

HUM (T0 )
HUv (T0 ) = (23.303)
vM

respectively
H0M (T0 )
H0v (T0 ) = (23.304)
vM

3
Note that the standard volume32 is vM = 22.41 kmol
m
.

32 At standard conditions, i.e. at 101.325 kPa and 0 ◦ C.


23.4 Energetic Balancing 779

23.4.6 Combustion Temperature

(I.) Combustion Temperature Above Dew Point ϑτ

In Sect. 23.4.2, the following energy balance has been derived for combustion with
unsaturated humid air, see Eq. 23.257. In case the combustion temperature TEG is
above the dew point,33 i.e. no condensation occurs,34 it is

q = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +


 
− [h F (TF ) − h F (T0 )] − λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) − HU (T0 ).
(23.305)
With the following abbreviations

h in (ϑa , ϑF , λ) = [h F (TF ) − h F (T0 )] +


  (23.306)
+ λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) + HU (T0 )

and

h out (ϑEG , λ) = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )]


(23.307)
the specific released heat −q is:

−q = h in (ϑa , ϑF , λ) − h out (ϑEG , λ) (23.308)

(II.) Combustion Temperature Below Dew Point ϑτ

In Sect. 23.4.2 the following energy balance has been derived for combustion with
unsaturated humid air, see Eq. 23.260, assuming that the combustion temperature
TEG is below the dew point, i.e. the vapour partially condenses35 :

q = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +


 
− [h F (TF ) − h F (T0 )] − λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) − HU (T0 )+
ṁ liq
− h v (TEG ) .
ṁ F
(23.309)
With the following abbreviations

33 No step 4 is required.
34 In this example even no condensation takes place in the pre-conditioning in step 1.
35 In this case, a step 4 is required. In this example, however, no condensation may occur during

preconditioning in step 1.
780 23 Combustion Processes
 
h in (ϑa , ϑF , λ) = [h F (TF ) − h F (T0 )] + λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) +
+ HU (T0 )
(23.310)
and

h out (ϑEG , λ) = μEG · [h EG (TEG ) − h EG (T0 )] +


ṁ liq (23.311)
+ a · [h Ash (TEG ) − h Ash (T0 )] − h v (TEG )
ṁ F

the specific released heat −q is:

−q = h in (ϑa , ϑF , λ) − h out (ϑEG , λ) (23.312)

The mass of the condensing water ṁ liq has been investigated in Sect. 23.3.6. It is

M H2 O
ṁ liq μEG,H2 O − xs μEG
= μ∗EG,H2 O =
MEG
(23.313)
ṁ F 1 − xs

where xs is a function of the total pressure p and the saturation pressure ps as a


function of the exhaust gas temperature, see also Eq. 23.144,

ps (TEG )
xs = . (23.314)
p

The lower the exhaust gas temperature, the more water condenses.

The h, ϑ-Diagram

The equation derived in the previous sections36

h in (ϑa , ϑF , λ) − h out (ϑEG , λ) = −q > 0 (23.315)

can be visualised in a so-called h, ϑ-diagram, which is exemplified in Fig. 23.20.


Enthalpy differences of the fuel can usually be neglected compared to the lower
heating value, so that h in simplifies to
 
h in (ϑa , ϑF , λ) ≈ h in (ϑa , λ) = λlmin · h 1+x (ϑa ) − h 1+x (ϑ0 ) + HU (ϑ0 ). (23.316)

The functions h in and h out can be calculated and plotted depending on the temperature,
with the fuel-air equivalence ratio and the humidity as parameters. A combustion,
shown in a h, ϑ-diagram, see Fig. 23.20, starts exemplarily at (Start), i.e. with air
being supplied with

36 This equation counts for combustion temperatures below and above the dew point.
23.4 Energetic Balancing 781

Fig. 23.20 h, ϑ-diagram

ϑa = ϑF = ϑ0 . (23.317)

This determines the enthalpy h in supplied to the system. In case (1), a specific heat of
q < 0 is released during combustion and the exhaust gas temperature is ϑEG,1 . The
remaining enthalpy is released with the exhaust gas and ash. If there is no ash, it is

h out (ϑEG , λ) = μEG · [h EG (ϑEG ) − h EG (ϑ0 )] = h EG . (23.318)

Obviously, the exhaust gas temperature is above the dew point so that no water
condenses. The higher the exhaust gas temperature is, the less heat is released during
combustion. Finally, in case (2), no heat is released during combustion and the
exhaust gas temperature is at a maximum. This temperature is called adiabatic flame
temperature ϑEG, adiabat , and the entire energy is released as exhaust gas/ash enthalpy.
At the dew point, the h out curve bends. A comparison of Eqs. 23.307 and 23.311

shows that this is due to the condensation of water, i.e. ṁliqF h v (ϑEG ) which occurs
when the temperature falls below the dew point.
According to Fig. 23.20, in case the combustion runs isothermally at reference
temperature ϑ0 , with the entire water being hypothetically in vapour state, the released
specific heat is HU (ϑ0 ), i.e. it fulfils the definition of the lower heating value, see Sect.
23.4.1. However, if the entire product water under these conditions condenses and
liquefies, the released heat would be increased by the specific enthalpy of vaporisation
μEG,H2 O h V (ϑ0 ), see Fig. 23.20. By definition, see Sect. 23.4.3, this released heat is
the specific higher heating value. Realistically, according to the p, T -diagram, see
Fig. 18.8, not all of the water can be condensed, but depending on the exhaust gas
782 23 Combustion Processes

temperature, some of the water is still present in vapour state. Thus, only the scaled
specific enthalpy of vaporisation μ∗EG,H2 O h V (ϑ0 ) is released. This is represented by
point (V) in Fig. 23.20. According to this consideration, the maximum specific heat
that can be released is smaller than the specific higher heating value and is marked
in the h, ϑ-diagram as qmax .

Calculation of the Exhaust Gas Temperature

Now that it is known how the dependency between released heat and exhaust gas
temperature can be determined graphically, the exhaust gas temperature is calculated
in a next step. Since the thermal properties of the exhaust gas are temperature-
dependent, numerical methods are needed to solve the correlation between released
heat and exhaust gas temperature. This is shown for two cases:
• Adiabatic flame temperature (dry air)

If the combustion air is dry, see Fig. 23.21, the first law of thermodynamics follows
Eq. 23.240 under the premise that the adiabatic flame temperature is above dew
point. Since combustion is supposed to be adiabatic, the specific heat is q = 0.
Accordingly, the first law of thermodynamics states

0 = − [h F (TF ) − h F (T0 )] − λlmin · [h a (Ta ) − h a (T0 )] − HU (T0 )+



+ μEG,i · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (Tad ) − h Ash (T0 )] .
i
(23.319)
The application of the caloric equations of state for the specific enthalpies requires
averaged specific heat capacities, see Eqs. 23.247–23.252. Thus, the first law of
thermodynamics can be solved for ϑad , i.e.

HU (T0 ) + c p,F |ϑϑF0 · (ϑF − ϑ0 ) + λ · lmin · c p,a |ϑϑa0 · (ϑa − ϑ0 )


ϑad = ϑ0 +
μEG · c p,EG |ϑϑad0 + a · cAsh |ϑϑad0
(23.320)

Fig. 23.21 Adiabatic flame


temperature (dry air)
23.4 Energetic Balancing 783

Unfortunately, this is an implicit equation, since the specific heat capacities of


the exhaust gas and ashes also depend on the adiabatic flame temperature ϑad .
Therefore, an iterative, numerical solution strategy needs to be followed:
(1) Start with an initial temperature ϑad,0
(2) Calculate specific heat capacities for exhaust gas and ashes at this temperature
(3) Calculate ϑad,n+1 following Eq. 23.320
(4) Repeat steps 2 and 3 until the deviation (ϑad,n+1 − ϑad,n ) is sufficiently small
Alternatively, the adiabatic flame temperature can be determined graphically in a
h, ϑ-diagram, see Fig. 23.20.

• Exhaust gas temperature—Generic approach

A generic combustion process37 follows Fig. 23.16. Humid air is used, which
additionally carries liquid water into the combustion chamber. The first law of
thermodynamics has been derived with Eq. 23.266 and thus obeys

q = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] +


 
− [h F (TF ) − h F (T0 )] − λlmin h 1+x (Ta , x) − h 1+x (T0 , x) +
ṁ liq (23.321)
− HU (T0 ) + λlmin xliq (T0 ) h v (T0 ) − h v (TEG ) .

 F  
for TEG <Tτ

Note that the heat of vaporisation with the last term only occurs when the com-
bustion temperature falls below the dew point, which depends on the exhaust gas
composition, cf. Sect. 23.3.6. The function q = f (TEG ) can be easily calculated,
taking into account the temperature dependence of the thermal properties and
clarifying whether water eventually condenses. However, the inverse function

TEG = f (q) (23.322)

is rather complex, since such an equation can not be solved explicitly for TEG . This
is because
– the specific enthalpies of exhaust gas,
– the specific enthalpies of ashes and, in case water condenses in step 5,
– the specific heat of vaporisation
depend on the exhaust gas temperature. Therefore, numerical methods must be
applied to solve Eq. 23.322.

37Under the premise, that no condensation takes place in step 1 when air is pre-conditioned from
Ta to T0 .
784 23 Combustion Processes

Fig. 23.22 Combustion


chamber

23.5 Combustion Chamber

A combustion chamber is shown exemplary in Fig. 23.22. Fuel as well as air are
supplied, exhaust gas and ashes38 are released. The utilised specific heat shall be
qE = ṁQ̇FE . However, due to insulation losses, the specific heat release qiso = Q̇ṁisoF
cannot be used further technically. The corresponding energy balance39 obeys
 
qE + qiso = − [h F (TF ) − h F (T0 )] − λlmin · h 1+x (Ta ) − h 1+x (T0 ) − HU (T0 )+

+ μEG,i · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )] .
i
(23.323)

23.5.1 Efficiency

The efficiency of combustion is defined with the usable thermal energy according to

|qE |
ηK = . (23.324)
HU

However, the overall combustion efficiency follows

|qE + qiso |
ηF = . (23.325)
HU

The specific enthalpy released with the exhaust gas is given by

38In case of solid fuels.


39If the combustion air is moist but unsaturated before and after step 1 and if the combustion
temperature is above the dew point.
23.5 Combustion Chamber 785

h EG = μEG,i · [h EG (EG ) − h EG (T0 )] (23.326)
i

with
μEG,i = μEG = (1 − a) + λ · lmin · (1 + x). (23.327)
i

To maximise the heat output of the combustion chamber, the exhaust gas temperatures
should be low.

23.5.2 Operation

In this section, the influence of air preheating as well as the fuel-air equivalence ratio
λ on the combustion temperature is discussed. Figure 23.23, which is based on the
h, ϑ-diagram introduced earlier, shows that when the air is preheated, the supplied
specific enthalpy h in increases, i.e.

h in (ϑa , ϑF , λ) = [h F (TF ) − h F (T0 )] +


  (23.328)
+ λlmin · h 1+x (Ta , x) − h 1+x (T0 , x) + HU (T0 ).

Consequently, the adiabatic flame temperature rises. Figure 23.23 also shows that
the adiabatic flame temperature decreases when the fuel-air equivalence ratio λ is
increased. This is caused by the increasing mass of the reactants. The chemically
bonded energy of the fuel is required to heat a larger mass.40 This leads to a lower
temperature rise. The supplied specific enthalpy h in , according to Eq. 23.328, as well
as the released specific enthalpy h out

h out (ϑEG , λ) = μEG · [h EG (TEG ) − h EG (T0 )] + a · [h Ash (TEG ) − h Ash (T0 )]


(23.329)
run with a larger gradient with increasing λ. Figure 23.24 shows the influence of the
two parameters exemplarily for the combustion of oil: Preheating the combustion air
increases the exhaust gas temperature, increasing the air volume lowers it.

Problem 23.5 A fuel cell is operated stoichiometrically with hydrogen and oxygen.
Both reactants are supplied with ϑ = 25 ◦ C, the inlet pressure of the hydrogen is
pF = 200 kPa. The product water is released in liquid state at ϑW = 60 ◦ C with a
specific heat capacity in liquid state of cW = 4.19 kgkJK . The fuel cell generates a
DC-voltage of Uel = 52 V and a current of Iel = 118 A with a fuel consumption of
3
VF = 2.2 mh . The specific lower heating value of the fuel at a reference temperature

40 Especially the excess oxygen and the inert nitrogen consume sensible heat.
786 23 Combustion Processes

Fig. 23.23 Combustion chamber—variation of λ

Fig. 23.24 Influence of fuel-air equivalence ratio and air preheating on the adiabatic flame tem-
perature for oil according to [7]

of ϑ0 = 25 ◦ C is HU = 119.946 MJ
kg
. Water has a specific heat of vaporisation at 25 ◦ C
of h V = 2442 kg kJ
.
(a) Calculate the mass fluxes of the fuel ṁ F , of the required oxygen ṁ O2 and of the
product water ṁ W .
(b) What heat flux Q̇ needs to be released by the fuel cell?
(c) Calculate the efficiency of the fuel cell, which is defined as η = m−P el
˙F HU
.
23.5 Combustion Chamber 787

Solution

(a) Hydrogen is treated as an ideal gas so that the mass flux of the fuel ṁ F follows
the thermal equation of state

pF V̇F = ṁ F RF T. (23.330)

The individual gas constant is

RM J
RF = = 4157.1 . (23.331)
MF kg K

Hence, the mass flux of hydrogen is

pF V̇F kg
ṁ F = = 9.8610 × 10−5 . (23.332)
RF T s

Since the fuel is pure hydrogen, its elemental analysis yields h = 1. The mini-
mum oxygen demand under stoichiometric conditions is therefore given by

omin = 2.667c + 8h = 8. (23.333)

The mass flux of the oxygen follows accordingly

kg
ṁ O2 = omin ṁ F = 7.8888 × 10−4 . (23.334)
s
Hydrogen and oxygen react and water occurs with

kg
ṁ W = ṁ O2 + ṁ F = 8.8749 × 10−4 . (23.335)
s
(b) The energy balance follows Fig. 23.25 and is split into three successive steps.

• Step 1—Reaction

Preconditioning is not necessary in this case, as hydrogen and oxygen are


supplied at reference temperature. In step 1, the (electrochemical) reaction is
carried out so that all the chemically bonded energy is released as thermal
energy and electrical energy, i.e.

Q̇ 1 + Pel = −ṁ F HU . (23.336)

The released electrical power follows

Pel = −Uel Iel . (23.337)


788 23 Combustion Processes

Fig. 23.25 Solution to Problem 23.5b

Hence, the released heat is

Q̇ 1 = −ṁ F HU + Uel Iel = −5.6918 kW. (23.338)

Note that the product water leaving step 1 is in a gaseous state as the lower
heating value has been applied.
• Step 2—Latent heat

The first law of thermodynamics for step 2 obeys

Q̇ 2 + ṁ W h v (T0 ) = ṁ W h liq (T0 ). (23.339)

The released heat flux follows


 
Q̇ 2 = ṁ W h liq (T0 ) − h v (T0 ) = −ṁ W h v (T0 ) = −2.1672 kW. (23.340)

• Step 3—Sensible heat

In step 3, the liquid water is postconditioned from 25 to 60 ◦ C. Since there is


no phase change, it is pure sensible heat, and the first law of thermodynamics41
yields
Q̇ 3 + ṁ W h liq (T0 ) = ṁ W h liq (TW ). (23.341)

The solution for the heat flux and the application of the caloric equation of
state for incompressible liquids leads to
 
Q̇ 3 = ṁ W h liq (TW ) − h liq (T0 ) = ṁ W cW (TW − T0 ) = 0.1302 kW.
(23.342)

41 Step 3 is supposed to be isobar.


23.5 Combustion Chamber 789

The overall thermal energy flux is

Q̇ = Q̇ 1 + Q̇ 2 + Q̇ 3 = −7.7289 kW. (23.343)

(c) The efficiency of the fuel cell follows

−Pel
η= = 0.5188. (23.344)
m˙F HU

Thus, 51.88% of the chemically bonded energy is converted into electricity. The
rest is released as heat.

Problem 23.6 A fossil fuel (c = 0.85, h = 0.15) is fired with air composed of O2 ,
N2 and CO2 , see Table 23.7. The mass flux of the air is ṁ a = 8 kgs .
(a) What is the maximum mass flux of fossil fuel that can be oxidised?
(b) Calculate the mass concentrations of the exhaust gas ξEG,i .
(c) Calculate the adiabatic flame temperature when the air is supplied at 100 ◦ C and
the fuel at 25 ◦ C. The lower heating value of the fuel is HU (25 ◦ C) = 42.6 MJ
kg
.
The required averaged specific heat capacities can be taken from Table 23.8.

Solution

(a) The minimum oxygen demand is

Table 23.7 Composition of the combustion air


O2 N2 CO2
ξi 0.12 0.70 0.18

Table 23.8 Averaged specific heat capacities c p |ϑ0 in kJ


kg K
ϑ [◦ C] O2 N2 CO2 H2 O
0 0.9147 1.0394 0.8173 1.8589
25 0.9163 1.0395 0.8307 1.8615
100 0.9229 1.0404 0.8684 1.8724
200 0.9354 1.0435 0.9128 1.8931
1100 1.0434 1.1254 1.1413 2.1744
1200 1.0511 1.1343 1.1560 2.2075
1300 1.0583 1.1426 1.1693 2.2399
790 23 Combustion Processes

ṁ O2 ,min
omin = = 2.667c + 8h + s − o. (23.345)
ṁ F

The available oxygen mass flow is given by the concentration of the combustion
air, i.e.
ṁ O2 ,min = ξO2 ṁ a . (23.346)

A combination of both equations leads to

ξO2 ṁ a kg
ṁ F = = 0.2769 . (23.347)
2.667c + 8h + s − o s

(b) The composition of the exhaust gas μEG,i is


• Carbon monoxide

ξCO2 ṁ a
μEG,CO2 = 3.667c
   + = 8.3174 (23.348)

comb.  F 
air

• Water
μEG,H2 O = 9h = 1.35 (23.349)

• Oxygen
μEG,O2 = 0 (23.350)

There is no oxygen in the exhaust gas42 because λ = 1.


• Nitrogen
ξN ṁ a
μEG,N2 = 2 = 20.2239 (23.351)
ṁ F

For the exhaust gas this results in



μEG = μEG,i = 29.8912. (23.352)
i

Accordingly, the concentrations are:


• Carbon monoxide
μEG,CO2
ξEG,CO2 = = 0.2783 (23.353)
μEG

• Water
μEG,H2 O
ξEG,H2 O = = 0.0452 (23.354)
μEG

42 The maximum mass flux of fuel in (a) has been asked for.
23.5 Combustion Chamber 791

• Oxygen
μEG,O2
ξEG,O2 = =0 (23.355)
μEG

• Nitrogen
μEG,N2
ξEG,N2 = = 0.6766 (23.356)
μEG

(c) The equation for the adiabatic flame temperature has been derived in Sect. 23.4.6
and follows, see Eq. 23.320,

HU (T0 ) + c p,F |ϑϑF0 · (ϑF − ϑ0 ) + λ · lmin · c p,a |ϑϑa0 · (ϑa − ϑ0 )


ϑad = ϑ0 + .
μEG · c p,EG |ϑϑad0 + a · cAsh |ϑϑad0
(23.357)
The fuel is supplied at reference temperature, no ash is produced and it is a
stoichiometric combustion with λ = 1, so that the equation simplifies to

HU (T0 ) + l · c p,a |ϑϑa0 · (ϑa − ϑ0 )


ϑad = ϑ0 + (23.358)
μEG · c p,EG |ϑϑad0

with
ṁ a
l= = 28.8912 = μEG − 1. (23.359)
ṁ F

According to Eq. 12.191 the averaged specific heat capacity is

1  
c p,EG |ϑϑad0 = c p,EG |ϑ0 ad ϑad − c p,EG |ϑ0 0 ϑ0 (23.360)
ϑad − ϑ0

with
c p,EG = ξEG,i c p,EG,i . (23.361)

For the air it follows accordingly

1   kJ
c p,a |ϑϑa0 = c p,a |ϑ0 a ϑa − c p,a |ϑ0 0 ϑ0 = 0.9981 (23.362)
ϑa − ϑ0 kg K

with
c p,a = ξi c p,i . (23.363)

The averaged specific heat capacities can be taken from Table 23.8. Obvi-
ously, Eq. 23.357 can be solved numerically, see the iterations documented in
Table 23.9.
792 23 Combustion Processes

Table 23.9 Iterations


Iteration ϑad,i in ◦ C c p,EG |ϑϑad
0
in kJ
kg K ϑad,i+1 in ◦ C
0 1300 1.2031 1269.7
1 1269.7 1.1999 1273.0
2 1273.0 1.2002 1272.7
3 1272.7 1.2002 1272.7

After 3 iterations, the solution does not change so that the adiabatic flame temperature
is ϑad = 1272.7 ◦ C.

Problem 23.7 In a combustion chamber methane (CH4 ) and propane (C3 H8 ) are
fired isobarically with a fuel-air equivalence ratio of λ = 1.3, see Fig. 23.26. Both
fuels are supplied with a temperature of 25 ◦ C and a pressure of 1 bar. The humid
combustion air has a temperature of 25 ◦ C and a relative humidity of ϕ = 60%. The
mass fractions of the dry air are ξO2 ,dry = 0.23 and ξN2 ,dry = 0.77. A heat flux of Q̇ D
is released, so that the exhaust gas temperature is 200 ◦ C. The pressure shall be 1 bar.
It further is
• Methane: V̇CH4 = 0.5 mh , lower heating value HU,CH4 (25 ◦ C) = 50.01 MJ
3

kg
• Propane: V̇C3 H8 = 0.3 mh , lower heating value HU,C3 H8 (25 ◦ C) = 46.35 MJ
3

kg
• Environment: ϑenv = 25 ◦ C, p = 1 bar
Insulation losses and changes in kinetic and potential energies can be neglected.
(a) Calculate the mass fluxes of methane and propane. Both shall be treated as ideal
gases.
(b) Determine the total mass fractions of the mixture of methane and propane c and
h.
(c) What is the minimum oxygen demand omin ? Calculate the mass flux of the humid
air.

Fig. 23.26 Sketch to Problem 23.7


23.5 Combustion Chamber 793

(d) Calculate the concentrations of the exhaust gas ξEG,i .


(e) What is the dew point of the exhaust gas?
(f) What is the overall lower heating value HU,F (25 ◦ C) of the fuel mixture?
(g) Calculate the released heat flux Q̇ D ?

Solution

(a) The mass fluxes result from the thermal equation of state, since both gases are
to be treated as ideal gases, i.e.

p V̇
ṁ = . (23.364)
RT
The individual gas constants can be determined with the molar masses

kg
MCH4 = 16 (23.365)
kmol
and
kg
MC3 H8 = 44 . (23.366)
kmol
So, the mass fluxes follow

p V̇CH4 MCH4 kg
ṁ CH4 = = 8.9645 × 10−5 (23.367)
RM T s

respectively

p V̇C3 H8 MC3 H8 kg
ṁ C3 H8 = = 1.4791 × 10−4 . (23.368)
RM T s

(b) The individual mass fractions of the fuels are


• Methane
1MC
cCH4 = = 0.75 (23.369)
MCH4

and
4MH
h CH4 = = 0.25 (23.370)
MCH4

• Propane
3MC
cC3 H8 = = 0.8182 (23.371)
MC3 H8
794 23 Combustion Processes

and
8MH
h C3 H8 = = 0.1818 (23.372)
MC3 H8

The following therefore applies to the mixture of both fuels

ṁ CH4 cCH4 + ṁ C3 H8 cC3 H8


c= = 0.7925 (23.373)
ṁ CH4 + ṁ C3 H8

and
ṁ CH4 h CH4 + ṁ C3 H8 h C3 H8
h= = 0.2075. (23.374)
ṁ CH4 + ṁ C3 H8

(c) The minimum oxygen demand is

omin = 2.667c + 8h = 3.7738. (23.375)

The minimum dry air demand is


omin
lmin = = 16.408. (23.376)
ξO2 ,dry

The overall demand of dry air obeys

l = λlmin = 21.33. (23.377)

The mass flux of dry air follows



ṁ a = l ṁ CH4 + ṁ C3 H8 . (23.378)
  
=ṁ F

To determine the mass flux of humid air, the vapour content x of the combustion
air is required,43 i.e.

ϕ · ps (25 ◦ C)
x = 0.622 = 0.0121. (23.379)
p − ϕ · ps (25 ◦ C)

Hence, the mass flux of the humid air is

kg
ṁ A = (1 + x)ṁ a = (1 + x)l ṁ CH4 + ṁ C3 H8 = 18.4621 . (23.380)
h
(d) The composition of the exhaust gas μEG,i is
• Carbon monoxide

43 According to the steam table it is ps (25 ◦ C) = 0.0317 bar.


23.5 Combustion Chamber 795

μEG,CO2 = 3.667c = 2.9059 (23.381)

• Water

μEG,H2 O = 9h + λlmin x = 2.1252 (23.382)

• Oxygen

μEG,O2 = omin (λ − 1) = 1.1322 (23.383)

• Nitrogen

μEG,N2 = λlmin ξN2 ,dry = 16.4244 (23.384)

For the exhaust gas this results in



μEG = μEG,i = 22.5877. (23.385)
i

The concentrations are therefore:


• Carbon monoxide
μEG,CO2
ξEG,CO2 = = 0.1287 (23.386)
μEG

• Water
μEG,H2 O
ξEG,H2 O = = 0.0941 (23.387)
μEG

• Oxygen
μEG,O2
ξEG,O2 = = 0.0501 (23.388)
μEG

• Nitrogen
μEG,N2
ξEG,N2 = = 0.7271 (23.389)
μEG

(e) In case condensation occurs, it is

ps (ϑτ )
xEG,H2 O = . (23.390)
p
796 23 Combustion Processes

Therefore, the mass fraction of the water must be converted into molar fractions,
i.e.
MEG
xEG,H2 O = ξEG,H2 O (23.391)
MH2 O

with
RM
MEG = . (23.392)
REG

The gas constant of the exhaust gas is


RM J
REG = ξRG,i Ri = ξRG,i = 296.7 . (23.393)
i i
Mi kg K

The molar mass is therefore


RM kg
MEG = = 28.02 . (23.394)
REG kmol

The mass fraction of the product water is

MEG
xEG,H2 O = ξEG,H2 O = 0.1465. (23.395)
MH2 O

The critical partial pressure obeys

ps (ϑτ ) = xEG,H2 O p = 0.1465 bar. (23.396)

According to the steam table, the dew point temperature is finally

ϑτ = 53.48 ◦ C. (23.397)

(f) The overall lower heating value HU,F (25 ◦ C) of the fuel mixture obeys

!
ṁ F HU,F (25 ◦ C) = ṁ CH4 HU,CH4 (25 ◦ C) + ṁ C3 H8 HU,C3 H8 (25 ◦ C). (23.398)

Solving for HU,F (25 ◦ C) brings

ṁ CH4 HU,CH4 (25 ◦ C) + ṁ C3 H8 HU,C3 H8 (25 ◦ C) MJ


HU,F (25 ◦ C) = = 47.731 .
ṁ CH4 + ṁ C3 H8 kg
(23.399)
(g) The heat flux for the combustion with unsaturated humid air is,44 cf. Eq. 23.256,

44 The exhaust gas temperature is above the dew point.


23.5 Combustion Chamber 797

Q̇ D =ṁ EG · [h EG (TEG ) − h EG (T0 )] + ṁ Ash · [h Ash (TEG ) − h Ash (T0 )] +


 
− ṁ F [h F (TF ) − h F (T0 )] − ṁ a · h 1+x (Ta , x) − h 1+x (T0 , x) +
− ṁ F HU,F (T0 ).
(23.400)
Fuel and air are supplied at reference temperature and no ash is produced, so the
energy balance is simplified as follows

Q̇ D = −ṁ F HU,F (T0 ) + ṁ EG · [h EG (TEG ) − h EG (T0 )] . (23.401)

The application of the caloric equation of state for the exhaust gas leads to
 ◦ ◦ 
Q̇ D = −ṁ F HU,F (T0 ) + ṁ EG · c p,EG |200
0
C
200 ◦ C − c p,EG |25
0
C
25 ◦ C
(23.402)
with

◦ kJ
c p,EG |200
0
C
= ξEG,i c p,i |200
0
C
= 1.1004 (23.403)
i
kg K

and

◦ kJ
c p,EG |25
0
C
= ξEG,i c p,i |25
0
C
= 1.0837 . (23.404)
i
kg K

Finally, the heat flux is


 ◦ ◦ 
Q̇ D = −ṁ F HU,F (T0 ) + ṁ EG · c p,EG |200
0
C
200 ◦ C − c p,EG |25
0
C
25 ◦ C
= −10.3 kW.
(23.405)
Chapter 24
Chemical Reactions

In this chapter, chemical reactions are studied more generally than in the previ-
ous chapters. In Chap. 23, the focus has been on fossil fuel combustion and the
lower/higher heating value has been introduced to treat a chemical conversion in
terms of energy. However, this method used to be rather impractical as the oxidation
has to be divided into several parts and it has been necessary to distinguish whether
condensation takes place. Now a simpler method is preferred by introducing the
so-called absolute specific enthalpy/entropy. Here, the specific enthalpy refers not
only to a caloric effect, as has been the case in Parts I and II, but also to the specific
chemical binding energy: When determining the specific enthalpies, it is necessary to
consider that the composition of the mass fluxes involved is different, since chemical
components are formed and consumed. The calculation of enthalpy differences must
therefore be based on mutually consistent absolute values of the enthalpies.

24.1 Mass Balance

For a chemical reaction, the mass balances are first derived generically, i.e. it is
determined how much of each reactant is needed to obtain a certain amount of
products. This step has also been carried out recently in Chap. 23. However, this first
requires a chemical reaction scheme, i.e.

γA A + γB B → γC C + γD D. (24.1)

Components A and B react chemically and eventually lead to products C and D. The
stoichiometric factors γi actually show how particles, i.e. atoms or molecules, are
consumed and rearranged by the chemical reaction. The stoichiometric factors can
be determined on the premise that the number of any atom on the reactant side must
be balanced by the number of the same atom on the product side. According to the
© Springer Nature Switzerland AG 2022 799
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2_24
800 24 Chemical Reactions

rule of three, it can be easily linked to the molar quantity n i . The reaction schemes
can finally be referred to 1 particle of component A, i.e.
γB γC γD
1+ → + . (24.2)
γA γA γA

Multiplication by n A -particles then yields

γB γC γD
nA + nA → nA + nA (24.3)
γA γA γA

If a molar amount of n A particles of A is to be transformed by a chemical reaction


with B, the γγAB -fold of B is needed and the γγAC -fold of C is produced. With the molar
mass of the individual components
mi
Mi = (24.4)
ni

the converted masses can be easily calculated, i.e.

γB γC γD
n A MA + n A MB → n A MC + n A MD (24.5)
γA γA γA

Analogous to the previous chapter, the molar specific conversion rate can be defined
for component A, for example, i.e.

ni
νi = (24.6)
nA

The mass specific conversion rate follows accordingly

mi
μi = (24.7)
mA

This can be applied to the reactants


 
γB molB
νB = (24.8)
γA molA

and  
γB M B kgB
μB = . (24.9)
γA M A kgA
24.1 Mass Balance 801

For the products it is  


γC molC
νC = (24.10)
γA molA

and  
γC M C kgC
μC = (24.11)
γA M A kgA

respectively  
γD molD
νD = (24.12)
γA molA

and  
γD M D kgD
μD = . (24.13)
γA M A kgA

The product composition (“Pr”) can be given by the molar or mass concentration,
i.e.
n Pr,i
xPr,i =  (24.14)
n Pr,i

and
m Pr,i
ξPr,i =  (24.15)
m Pr,i

with 
xPr,i = 1 (24.16)
i

and 
ξPr,i = 1. (24.17)
i

According to the example shown, this results in


γC
xPr,C = (24.18)
γC + γD

and γD
xPr,D = (24.19)
γC + γD

respectively
γC M C
ξPr,C = (24.20)
γC M C + γD M D

and
802 24 Chemical Reactions

γD M D
ξPr,D = . (24.21)
γC M C + γD M D

Example 24.1 (Combustion of methanol with λ > 1) In this example methanol


CH3 OH shall be oxidised. The reaction scheme is now derived step by step:
• Stoichiometric with oxygen

In this case the reaction scheme follows


3
1 CH3 OH + O2 → 1 CO2 + 2 H2 O. (24.22)
  2
“F”

Note that the given stoichiometric coefficients satisfy the conservation of atoms:
The number of C-atoms on left and right side is 1, the H-atoms are 4 and the
O-atoms are 4 as well on both sides.

• Stoichiometric with air1

When supplied, air not only provides the required oxygen, but also nitrogen, which
must be included in the reaction scheme. As in the previous chapters, it shall be
treated as inert, i.e. it goes in and out without being chemically reactive, i.e.
3 79 3 79 3
1 CH3 OH + O2 + · N2 → 1 CO2 + 2 H2 O + · N2 . (24.23)
2 21 2 21 2
• Excess Air

A fuel-air equivalence ratio of λ > 1 results in excess air that is not consumed in
the reaction. Consequently, this unused air remains on the product side.
3 79 3
1 CH3 OH + λ · O2 + λ · · N2 → 1 CO2 + 2 H2 O
2 21 2
79 3 3
+ λ· · N2 + (λ − 1) · O2 . (24.24)
21 2 2
The fuel-air equivalence ration is defined as
n Air m Air
λ= = (24.25)
n Air, min m Air, min

see Sect. 23.3.1. Note that if λ = 1 the product side is free of oxygen.

Oxygen/Air Need
The molar oxygen demand can now be determined from the reaction scheme, i.e.

x N2 ,Air
1The ratio 79
21 is the molar ratio of atmospheric air x O2 ,Air . Thus, it shows how much more nitrogen
compared to oxygen is in the combustion air.
24.1 Mass Balance 803

3
n O2 = λ · · nF. (24.26)
2

In other words, λ · 23 -fold the amount of fuel is required as oxygen, see reaction
scheme Eq. 24.24. The molar specific oxygen demand is then

n O2 3
νO2 = ≡ O = λ · Omin = λ · (24.27)
nF 2

The mass consumption of oxygen follows accordingly with the molar masses

3 M O2
m O2 = λ · · mF (24.28)
2 MF

respectively in specific notation

m O2 3 MO2
μO2 = ≡ o = λ · omin = λ · (24.29)
mF 2 MF

In a next step the air demand can be calculated

3 1
na = λ · · nF · (24.30)
2 xO2 ,Air

respectively in specific notation

na 3 1
νa = ≡ L = λ · L min = λ · · (24.31)
nF 2 xO2 ,Air

The mass consumption of air follows accordingly with the molar masses

3 MO2 1
ma = λ · · mF · (24.32)
2 MF ξO2 ,Air

respectively in specific notation

ma 3 MO2 1
μa = ≡ l = λ · lmin = λ · · (24.33)
mF 2 MF ξO2 ,Air

Product Composition
The product composition is as follows and can be determined directly from the
chemical Eq. 24.24:
• Carbon dioxide
804 24 Chemical Reactions

n Pr,CO2 = 1 · n F (24.34)

Respectively in fuel-specific notation

n Pr,CO2
νPr,CO2 = =1 (24.35)
nF

• Water
n Pr,H2 O = 2 · n F (24.36)

Respectively in fuel-specific notation

n Pr,H2 O
νPr,H2 O = =2 (24.37)
nF

• Nitrogen
79 3
n Pr,N2 = λ · · · nF (24.38)
21 2
Respectively in fuel-specific notation

n Pr,N2 79 3
νPr,N2 = =λ· · (24.39)
nF 21 2

• Oxygen
3
n Pr,O2 = (λ − 1) · · nF (24.40)
2
Respectively in fuel-specific notation

n Pr,O2 3
νPr,O2 = = (λ − 1) · (24.41)
nF 2

24.2 Energy Balance

In Chap. 23 the energy balance of an oxidation with the specific lower heating value
HU (T0 ) of the fuel has been described. Fuel and air are supplied at a reference tem-
perature T0 , the exhaust gas leaves the system at the same reference temperature, and
the product water is completely in vapour state. Under these conditions, the specific
lower heating value represents the specific heat to be released in such an isother-
mal combustion. Although this has been a simple approach, several steps have been
required to fully describe the energetics of combustion and possible condensation.
24.2 Energy Balance 805

The specific enthalpy of each component involved covers the caloric part, i.e. the
specific sensible/latent heat influences the temperature respectively aggregate state of
each component. However, it has not been possible to include the chemically bonded
energy in the enthalpies. The reactants can therefore not be compared with each other
in terms of energy. Furthermore, it is not yet possible to take entropic effects into
account, so that the irreversibility of such combustion cannot be investigated.
In this section, a new approach is applied that extends the specific enthalpy to
include chemical effects and brings the enthalpies to an absolute reference level
that allows an energetic comparison of the reactants/products involved in a chemical
reaction. This is the key to performing energy balances such as those that have
been carried out in Parts I and II: In steady state, the flux of energy into an open
system is balanced by the flux of energy out of the system. A division into several
sub-steps, cf. Sect. 23.4, is then obsolete. Furthermore, the specific entropy is also
extended to chemical effects, so that an entropy balance for chemical reactions is
possible. However, these chemical effects require a uniform reference level for each
chemical substance. This new approach is based on the so-called specific absolute
enthalpy/entropy.

24.2.1 Caloric Equations of State

Enthalpy of Reaction/Formation, Absolute Enthalpy

The enthalpy of reaction 0R H is defined for standard conditions and describes the
change of enthalpy in isothermal/isobaric reactions, i.e. the difference between the
total enthalpies of reactants and products, see also [3]:

0R H ≡ HProducts
0
− HEducts
0
(24.42)

Since chemical equations describe how many particles react, e.g.

γA A + γB B → γC C + γD D (24.43)

molar specific enthalpies are applied. Therefore, the molar specific enthalpy2 of
reaction 0R Hm referred to reactant A obeys

γD γC γB
0R Hm = · Hm,D
0
+ · Hm,C
0
− · Hm,B
0
− Hm,A
0
(24.44)
γA γA γA

The standard conditions, although arbitrary, are usually defined as follows

2Though the absolute specific enthalpy is a specific state value, a capital letter Hm is used instead
of h M , to emphasise that the absolute specific enthalpy is meant. The same counts for the absolute
specific entropy Sm .
806 24 Chemical Reactions

T0 = 298.15 K (24.45)

and
p0 = 1 bar. (24.46)

0
This obviously requires the molar specific enthalpies Hm,i for each chemical sub-
stance. The following approach is now followed:
• The molar specific enthalpy of the elements i in their stable state at standard
conditions, e.g.
– hydrogen: H2 ,
– nitrogen: N2 ,
– oxygen: O2 ,
– carbon: C,
– sulphur: S2 ,
0
is arbitrarily set to Hm,i (T0 , p0 ) ≡ 0. This molar specific enthalpy is also denoted
absolute enthalpy at standard conditions.

• For each other chemical bond, the specific enthalpy of formation is determined
with reference to the standard conditions. These values can be determined from
experiments with a calorimeter, for example, or they have to be calculated from
already known values. For example, if the molar specific enthalpy of methane CH4
at standard conditions is to be known, the chemical equation is as follows

C(solid) + 2H2 → CH4 . (24.47)

Hence, the molar specific enthalpy of reaction at standard conditions yields

0R Hm = Hm,CH
0
4
− 2 · Hm,H
0
2
− Hm,C
0
= Hm,CH
0
4
. (24.48)

The standard molar specific enthalpy of this reaction 0R Hm provides the standard
enthalpy of formation 0B Hm,CH4 of methane. The requested absolute enthalpy of
0
methane Hm,CH4 is identical to the standard enthalpy of reaction in the formation
of methane from solid carbon and gaseous hydrogen, i.e.
0
Hm,CH4 = 0B Hm,CH4 = 0R Hm . (24.49)

For temperatures and pressures deviating from standard state, the caloric effect must
also be taken into account, so that the absolute specific enthalpy Hm,i is defined:
• Ideal gas

If the component is gaseous, it shall behave like an ideal gas, i.e. the absolute
specific enthalpy Hm,i (T ) is purely a function of temperature. With the molar
specific heat capacity C p , the following therefore applies to ideal gases:
24.2 Energy Balance 807

dHm,i = C p dT. (24.50)

An integration from standard state to a temperature T with the unified reference


level yields

T

Hm,i (T ) − Hm,i (T0 ) = C p

. (T − T0 ) .
0
(24.51)
  T0
=0B Hm,i (T0 )

Consequently, it is

T

Hm,i (T ) = 0B Hm,i (T0 ) + C p

(T − T0 ) . (24.52)
  T
chemical  0 
caloric
 
Listed in Appendix B.

The first term on the right-hand side of Eq. 24.52 represents the chemical energy
of formation and the second the caloric effect if the temperature deviates from
the standard temperature. For the absolute specific internal energy, proceed as
follows: The definition of enthalpy yields

Hm,i = Um,i + pvM,i . (24.53)

With the thermal equation of state, i.e.

pvM,i = RM T, (24.54)

the absolute specific internal energy follows

Um,i (T ) = Hm,i (T ) − RM T. (24.55)

• Incompressible liquid/solid

For an incompressible liquid/solid the absolute enthalpy3 is, cf. Eq. 12.132 in Part
I,
dHm,i = C dT + vM,i d p. (24.56)

An integration from standard state to a temperature T and pressure p with the


unified reference level yields

3 Note that there is no distinction between C p and Cv for incompressible liquids/solids.


808 24 Chemical Reactions

T

Hm,i (T, p) = 0B Hm,i (T0 , p0 ) + C

(T − T0 ) + vM,i ( p − p0 ) . (24.57)
  T  
chemical  0  caloric, p
caloric, T
 
Listed in Appendix B.

The last term of Eq. 24.57 can usually be neglected and is not included in Appendix
B for liquids. However, the following applies to the absolute specific internal
energy
Um,i (T, p) = Hm,i (T, p) − pvM,i . (24.58)

Hm,i is the so-called absolute specific enthalpy of a component i. It is listed for


several elements and bonds in the Appendix B. All components are supposed to be
ideal, i.e. they either follow an ideal gas or are incompressible liquids/solids.
Definition 24.1 The (specific molar) absolute enthalpies follow a uniform reference
system. For chemical compounds under standard conditions, they correspond to the
specific standard molar enthalpies of formation. These enthalpies of formation are
identical to the specific standard molar enthalpies of reaction. Deviating states can
be determined with the caloric equations of state of ideal fluids.

Absolute Entropy—Third Law of Thermodynamics

The Nernst heat theorem states that the entropy of a closed system for T → 0
approaches a constant that is independent of thermodynamic parameters, cf. [5]. This
means that the entropy change approaches zero when the temperature approaches
absolute zero, i.e.
lim S = 0. (24.59)
T →0

The theorem can be proven with the help of quantum mechanics. The third law of
thermodynamics, based on the heat theorem, states the following, see [3]:
Theorem 24.1 The entropy of any homogeneous, perfectly crystallised substance
approaches zero when the absolute temperature approaches zero Kelvin.
It also states that it is technically impossible to reach an absolute temperature of
0 K. The reference level T → 0 K for the entropy is necessary because substances at
T > T0 can occur in different states of aggregation and thus have different entropy
levels. Therefore, 0 K is an absolute reference level4 for any crystallised substance,
i.e. for a solid5 :
0
Sm,i (T0 = 0 K, p0 = 1 bar) = 0. (24.60)

4 Reference pressure p0 is the same as for the absolute enthalpy, whereas the reference temperature
is different.
5 Note that entropy is independent of pressure for incompressible solids.
24.2 Energy Balance 809

Since the entropy of all substances at absolute zero Kelvin is zero, there is no need
for a substance-dependent determination of the reference entropies as in the case
of the reference level for enthalpy. The specific molar entropy at constant pressure
p0 = 1 bar is obtained by integrating the entropy differential,6 starting at T0 :

T T
C p,i C p,i
Sm,i (T, p0 ) = 0
Sm,i (T0 = 0 K, p0 = 1 bar) + dT = dT (24.61)
  T T
=0 0 0

To achieve finite values, the following must apply7

lim C p,i (T, p) = 0. (24.62)


T →0

In Appendix B the absolute molar specific entropies for idealised components, i.e.
for ideal gases as well as for incompressible liquids, are listed:
• Ideal gas

For varying states T and pi the caloric equation of state, see Eq. 12.116, can be
applied. Note that if component i is part of a gaseous mixture, the partial pressure
pi of that component must be applied, see also Sect. 19.3.4, i.e.

T

T pi
Sm,i (T, pi ) = Sm,i
0
(T0 , p0 ) +C p,i

ln − RM ln (24.63)
  T0 T0 p0
=0

The temperature influence is already covered by the listed data in Appendix B.


Hence, the tables show Sm,i (T, p0 ), i.e.

T

T pi
Sm,i (T, pi ) = 0
Sm,i (T0 , p0 ) + C p,i

ln −RM ln . (24.64)
T0 T0 p 0
 
Sm,i (T, p0 )

Consequently, for varying pressures the following pressure correction is required


pi
Sm,i (T, pi ) = Sm,i (T, p0 ) −RM ln . (24.65)
  p0
Listed in Appendix B.

6 If it is a liquid/solid, it is Ci instead of C p,i .


7 The same applies to Cv,i .
810 24 Chemical Reactions

Fig. 24.1 Enthalpy of


reaction

• Incompressible liquid/solids

For incompressible liquids/solids, the absolute specific entropy purely is a function


of temperature, see Chap. 12. Hence, a pressure correction is not required and the
absolute specific entropy obeys

T

T
Sm,i (T ) = 0
Sm,i (T0 , p0 ) +Ci

ln = Sm,i (T, p0 ) . (24.66)


  T0 T0  
=0 Listed in Appendix B.

24.2.2 Open Systems

First Law of Thermodynamics

Since the absolute specific enthalpy Hm,i (T, p) now also includes the chemical
energy, the first law of thermodynamics can easily be treated for a chemical reaction,
cf. Fig. 24.1. In contrast to Sect. 23.4 no standardisation as in “Step 2” is required.
For a steady state process, the energy flux into the system must be equal to the energy
flux out of the system, i.e. the first law of thermodynamics states8
 
Q̇ + ṅ j Hm, j T j = ṅ i Hm,i (Ti ) (24.67)
j,Educts i,Products

respectively solved for the heat flux


 
Q̇ = ṅ i Hm,i (Ti ) − ṅ j Hm, j T j (24.68)
i,Products j,Educts

The absolute enthalpies Hm,i (T, p) can be found in Appendix B. Note that the data
listed are given as a function of T , so they can be applied directly to ideal gases. If
an incompressible liquid is involved, a correction of the absolute specific enthalpy
according to Eq. 24.57 is required.

8 Changes of kinetic/potential energies shall be ignored.


24.2 Energy Balance 811

Fig. 24.2 Enthalpy of


reaction at T0

If a reaction takes place under standard conditions, i.e. both products and reactants
have the reference temperature T0 , see Fig. 24.2, the absolute specific enthalpies
contain only the chemical energies. Thus, the first law of thermodynamics reads as

 
Q̇ = ṅ F 0R Hm (T0 ) = ṅ i 0B Hm,i (T0 ) − ṅ j 0B Hm, j (T0 ) (24.69)
i,Products j,Educts

0R Hm (T0 ) is called specific enthalpy of reaction at standard conditions. If the prod-
uct water is gaseous, the heat released corresponds to the lower heating value, i.e.

Q̇ = −ṅ F HUM (T0 ) . (24.70)

Thus, it is
 ṅ i 0  ṅ j
0R Hm (T0 ) = B Hm,i (T0 ) − 0B Hm, j (T0 )
i,Products
ṅ F j,Educts
ṅ F (24.71)
= −HUM (T0 ) .

Example 24.2 (Lower/higher heating value of methane) As an example the specific


enthalpy of reaction at standard conditions of methane is calculated. The chemical
equation follows
CH4 + 2O2 → CO2 + 2H2 O. (24.72)

According to Eq. 24.71 with the assumption that water occurs as vapour it is

0R Hm (T0 ) = 2 · 0B Hm,H2 O,g (T0 ) + 1 · 0B Hm,CO2 (T0 ) + (24.73)
−2· 0B Hm,O2 (T0 ) − 1 · 0B Hm,CH4 (T0 ) . (24.74)

With the absolute enthalpies taken from the tables in Appendix B it follows, that

kJ
0R Hm (T0 ) = −802.562 . (24.75)
mol
812 24 Chemical Reactions

Hence, following Eq. 24.71 the molar specific lower heating value is HUM =
kJ
802.562 mol . It can be converted into a mass specific lower heating value by using
kg
the molar mass of methane M = 16 kmol , i.e.

MJ
HU (25 ◦ C) = 50.1 (24.76)
kg

In case the product water is in liquid state it is

0R Hm (T0 ) = 2 · 0B Hm,H2 O,liq (T0 ) + 1 · 0B Hm,CO2 (T0 ) + (24.77)
−2· 0B Hm,O2 (T0 ) − 1 · 0B Hm,CH4 (T0 ) . (24.78)

With the absolute enthalpies taken from the tables in Appendix B it follows, that

kJ
0R Hm (T0 ) = −890.57 . (24.79)
mol

Hence, the molar specific higher heating value is: H0M = 890.57 mol kJ
. It can be con-
verted into the mass specific higher heating value by using the molar mass of methane
kg
M = 16 kmol , i.e.
MJ
H0 (25 ◦ C) = 55.6 (24.80)
kg

Example 24.3 (Theoretical adiabatic flame temperature of methane) The adiabatic


flame temperature of methane is to be determined as a function of the air-fuel equiv-
alence ratio λ > 1. Methane and air are supplied at ϑin = 25 ◦ C. To simplify, the
air is composed of 79 vol.−% nitrogen and 21 vol.−% oxygen. The stoichiometric
reaction equation for combustion with pure oxygen is as follows

CH4 + 2 O2 → CO2 + 2 H2 O. (24.81)

If combustion takes place with excess oxygen, the result is

CH4 + 2λ O2 → CO2 + 2 H2 O + 2 (λ − 1) O2 . (24.82)


79
However, combustion is carried out with air, which contains 21
times more nitrogen
than oxygen, so that

79 79
CH4 + 2λ O2 + 2λ N2 → CO2 + 2 H2 O + 2 (λ − 1) O2 + 2λ N2 . (24.83)
21 21
24.2 Energy Balance 813

If combustion is carried out as a steady state flow process, the first law of thermody-
namics yields9

ṅ CH4 Hm,CH4 (Tin ) + ṅ O2 Hm,O2 (Tin ) + ṅ N2 Hm,N2 (Tin ) + Q̇ + P =


 
=0
(24.84)
ṅ CO2 Hm,CO2 (Tad ) + ṅ O2 Hm,O2 (Tad ) +
+ ṅ N2 Hm,N2 (Tad ) + ṅ H2 O Hm,H2 O (Tad ) .

Division by the fuel molar flux ṅ CH4 and consideration of the stoichiometric factors
from the reaction equation results in

79
Hm,CH4 (Tin ) + 2λHm,O2 (Tin ) + 2λ Hm,N2 (Tin ) =
21
Hm,CO2 (Tad ) + 2 (λ − 1) Hm,O2 (Tad ) + (24.85)
79
+ 2λ Hm,N2 (Tad ) + 2Hm,H2 O (Tad ) .
21
The molar specific absolute enthalpies of the components involved can be taken
from the tables in Appendix B. Equation 24.85 therefore only has the adiabatic
flame temperature Tad as an unknown. The numerical solution of this equation is
shown in Fig. 24.3 as a function of the air-fuel equivalence ratio λ > 1.
It can be seen that the adiabatic flame temperature decreases with increasing air-fuel
equivalence ratio. This is due to an increase in the masses involved calorically in the
reaction, i.e. the chemical energy released must heat a larger mass. However, these are
theoretical values. In reality, so-called dissociation occurs at high temperatures above
1500 ◦ C, i.e. the reaction products decompose and the formation of NO occurs.10 This
dissociation is associated with energy consumption, so that realistic adiabatic flame
temperatures for small λ are lower than those shown in Fig. 24.3, see [3].

Second Law of Thermodynamics

A major disadvantage of the lower heating value approach is the impossibility of


determining the degree of irreversibility, i.e. applying the second law of thermody-
namics. With the method of absolute entropy, it is now possible to perform an entropy
balance and thus calculate the entropy generation. A steady state entropy balance for
a chemical reaction implies that the incoming entropy fluxes must be balanced by
the outgoing entropy fluxes. According to Fig. 24.4, the balance then obeys

Ṡa + Ṡi + ṠI = ṠII . (24.86)

9It is assumed that the water in the exhaust gas is present as vapour.
10 The chemical reaction pathways for dissociation are not applied in Eq. 24.85. Furthermore,
nitrogen is simplified considered inert.
814 24 Chemical Reactions

Fig. 24.3 Theoretical adiabatic flame temperature of methane

Fig. 24.4 Entropy and


irreversibility of a reaction

State (I) represents the inlet, i.e. the educts, state (II) represents the outlet, i.e. the
products. Since several educts respectively products may be present, the convective
entropy fluxes follow
 
ṠII − ṠI = ṅ i Sm,i (Ti , pi ) − ṅ j Sm, j T j , p j . (24.87)
i,Products j,Educts

Note that Ṡi represents the generation of entropy due to irreversibilities and Ṡa counts
for the entropy fluxes due to heat transfer. Thus, it is
 
Ṡi + Ṡa = ṅ i Sm,i (Ti , pi ) − ṅ j Sm, j T j , p j (24.88)
i,Products j,Educts
24.2 Energy Balance 815

Hence, for an adiabatic reaction it is Ṡa = 0. The molar specific entropy of an isother-
mal/isobaric reaction R Sm is defined as

!
 
ṅ F R Sm = ṠII − ṠI = ṅ i Sm,i (Ti , pi ) − ṅ j Sm, j T j , p j . (24.89)
i,Products j,Educts

 ṅ i  ṅ j
⇒ R Sm = Sm,i (Ti , pi ) − Sm, j T j , p j (24.90)

i,Products F

j,Educts F

According to this definition, the entropy balance for a steady state reaction is as
follows
ṅ F R Sm = Ṡa + Ṡi . (24.91)

24.2.3 Closed Systems

First Law of Thermodynamics

Let us now assume that the chemical reaction takes place in a closed system, see
Fig. 24.5: State (I) comprises the reactants that lead to the products in state (II).
State (I) and state (II) shall be in thermodynamic equilibrium, i.e. the educts have
a common temperature TI , the products a common temperature TII . The first law of
thermodynamics obeys

Q 12 + W12 = UII − UI + E a,I,II . (24.92)

For the change of internal energy it is

Fig. 24.5 First law of


thermodynamics reactive
closed system
816 24 Chemical Reactions
 
UII − UI = n i Um,i (TII ) − n j Um, j (TI ) . (24.93)
i,Products j,Educts

Since the specific absolute internal energies Um,i and Um, j are not listed in
Appendix B, they have to be calculated according to Eqs. 24.55 and 24.58 depending
on the aggregate state of the fluid:
• Ideal gas
Um,i (TII ) = Hm,i (TII ) − RM TII (24.94)

Um, j (TI ) = Hm, j (TI ) − RM TI (24.95)

• Incompressible liquid

Um,i (TII , pII ) = Hm,i (TII , pII ) − pII vM,i (24.96)

Um, j (TI , pI ) = Hm, j (TI , pI ) − pI vM, j (24.97)

Second Law of Thermodynamics

Figure 24.6 shows the entropy balance for a closed, reactive system. The second law
of thermodynamics obeys

Sa,12 + Si,12 = SII − SI . (24.98)

State (I) represents the educts, state (II) represents the products. Since several educts
respectively products may be present, the entropy balance follows
 
SII − SI = n i Sm,i TII , pII,i − n j Sm, j TI , pI, j . (24.99)
i,Products j,Educts

Fig. 24.6 Second law of


thermodynamics reactive
closed system
24.2 Energy Balance 817

The entropy being transferred with the heat is

II
δQ
Sa,12 = . (24.100)
T
I

Problem 24.1 In an electrolysis system ṁ W = 18 kgs water is chemically decom-


posed into hydrogen and oxygen under isothermal and isobaric conditions (ϑ =
25 ◦ C, p = 20 bar), see Fig. 24.7.
(a) What are the molar fluxes ṅ H2 O , ṅ H2 , ṅ O2 under steady state conditions?
(b) What are the enthalpy and entropy fluxes at inlet/outlet (water is supposed to be
liquid)?
(c) What is the required power Pt in reversible operation when heat is exchanged
with the environment (Tenv = 298.15 K)?

Solution

(a) The chemical reaction follows

2 H2 O → 2 H2 + 1 O2 . (24.101)

The molar flux of water is


ṁ H2 O kmol
ṅ H2 O = =1 . (24.102)
MH2 O s

Thus, according to the chemical equation it is

kmol
ṅ H2 = ṅ H2 O = 1 (24.103)
s
respectively

Fig. 24.7 Sketch to Problem


24.1
818 24 Chemical Reactions

1 kmol
ṅ O2 = ṅ H2 O = 0.5 . (24.104)
2 s
(b) The enthalpy and entropy fluxes are
• Inlet (I) @ 25 ◦ C, liquid

According to Appendix B it is

kJ
Hm,I,H2 O = −285.83 (24.105)
mol
and
kJ
Sm,I,H2 O = 69.942 . (24.106)
kmol K
The fluxes are

ḢI = ṅ H2 O Hm,I,H2 O = −285.83 × 103 kW (24.107)

and
kW
ṠI = ṅ H2 O Sm,I,H2 O = 69.942 . (24.108)
K
• Outlet (II)

At the outlet (II) oxygen and hydrogen occur. According to Appendix B this
results in
kJ
Hm,II,O2 = 0 (24.109)
mol
and
kJ
Hm,II,H2 = 0 . (24.110)
mol
The entire enthalpy flux is

ḢII = ṅ H2 Hm,II,H2 + ṅ O2 Hm,II,O2 = 0 kW. (24.111)

For the entropy it is

kJ kJ 20 kJ
Sm,II,O2 = 205.149 − 8.3143 ln = 180.24
kmol K kmol K 1 kmol K
(24.112)
and
kJ kJ 20 kJ
Sm,II,H2 = 130.681 − 8.3143 ln = 105.77 .
kmol K kmol K 1 kmol K
(24.113)
24.2 Energy Balance 819

The entire entropy flux is

kW
ṠII = ṅ H2 Sm,II,H2 + ṅ O2 Sm,II,O2 = 195.89 . (24.114)
K
(c) The first law of thermodynamics obeys

Q̇ + Pt,rev = ḢII − ḢI +  Ė a . (24.115)



=0

Obviously there are two unknowns ( Q̇ and Pt,rev ), so further information can be
obtained by the second law of thermodynamics, i.e.

Ṡa + Ṡi + ṠI = ṠII . (24.116)

The reaction shall be reversible, so that Ṡi = 0. Since heat is exchanged at a


constant ambient temperature, the following results


Ṡa = . (24.117)
Tenv

Hence, the heat flux is


Q̇ = Tenv ṠII − ṠI . (24.118)

Finally, the reversible technical power to be supplied is



Pt,rev = ḢII − ḢI − Tenv ṠII − ṠI = 248.27 MW. (24.119)

Problem 24.2 A fuel cell is operated stoichiometrically with hydrogen and oxy-
gen. Both reactants are supplied with ϑ = 25 ◦ C, the inlet pressure of hydrogen is
pF = 200 kPa. The product water is released in liquid state at ϑW = 60 ◦ C. The fuel
cell generates a DC-voltage of Uel = 52 V and a current of Iel = 118 A with a fuel
3
consumption of VF = 2.2 mh .

Note that this problem is identical with Problem 23.4. However, the problem is now
solved with the absolute enthalpy/entropy approach.
(a) Calculate the mass fluxes of the fuel ṁ F , of the required oxygen ṁ O2 and of the
product water ṁ W .
(b) What heat flux Q̇ needs to be released by the fuel cell?
820 24 Chemical Reactions

Solution

(a) Hydrogen is treated as an ideal gas, so that the mass fluxes of the fuel ṁ F follows
the thermal equation of state

pF V̇F = ṁ F RF T. (24.120)

The individual gas constant is

RM J
RF = = 4157.1 . (24.121)
MF kg K

Hence, the mass flux of hydrogen is

pF V̇F kg
ṁ F = = 9.8610 × 10−5 (24.122)
RF T s

Since the fuel is pure hydrogen, its elemental analysis is h = 1. The minimum
oxygen demand under stoichiometric conditions is therefore given by

omin = 2.667c + 8h = 8. (24.123)

The mass flux of the oxygen follows accordingly

kg
ṁ O2 = omin ṁ F = 7.8888 × 10−4 . (24.124)
s
Hydrogen and oxygen react and water is formed with

kg
ṁ W = ṁ O2 + ṁ F = 8.8749 × 10−4 . (24.125)
s
The molar fluxes are
ṁ W kmol
ṅ W = = 4.9305 × 10−5 (24.126)
MW s

and
ṁ F kmol
ṅ F = = 4.9305 × 10−5 (24.127)
MF s

and
ṁ O2 kmol
ṅ O2 = = 2.4652 × 10−5 . (24.128)
MO2 s
24.2 Energy Balance 821

(b) The first law of thermodynamics obeys

Pel + Q̇ + ḢI = ḢII (24.129)

with
Pel = −Uel Iel . (24.130)

Thus, the heat flux follows

Q̇ = ḢII − ḢI + Uel Iel . (24.131)

According to Appendix B the enthalpies are

ḢI = ṅ O2 Hm,O2 (25 ◦ C) + ṅ F Hm,H2 (25 ◦ C) = 0 (24.132)

and
kJ
ḢII = ṅ W Hm,H2 O,liq (60 ◦ C) = −283.1933 · ṅ W = −13.9628 kW.
mol
(24.133)
Hence, the heat flux finally is

Q̇ = ḢII − ḢI + Uel Iel = −7.8268 kW. (24.134)

24.3 Gibbs Energy

24.3.1 Definition

A new state value, the specific Gibbs energy, is defined as a combination of other
state values, i.e.
g = g (T, p) = h − T s = u + pv − T s (24.135)

It has already been introduced in Chap. 12, where the caloric state equations have
been derived. Its differential is thus given by

dg = dh − T ds − s dT = du + v d p + p dv − T ds − s dT (24.136)

Applying the fundamental equation of thermodynamics, see Eq. 13.31, it is

dg = dh − T ds − s dT = −s dT + v d p. (24.137)

However, this total differential must also follow


   
∂g ∂g
dg = dT + d p. (24.138)
∂T p ∂p T
822 24 Chemical Reactions

Fig. 24.8 Gibbs


energy—closed system

A comparison of Eqs. 24.137 and 24.138 leads to the change of Gibbs energy with
respect to temperature
 
∂g
= −s (24.139)
∂T p

and to the change of Gibbs energy with respect to pressure


 
∂g
=v (24.140)
∂p T

However, the specific Gibbs energy as state value of a fluid is constant, i.e. dg = 0,
for isobaric/isothermal conditions.

24.3.2 Molar Gibbs Energy

The extensive Gibbs energy follows from its definition, see Eq. 24.135, i.e.

G = mg (T, p) = H − T S. (24.141)

Replacing the mass with the molar quantity and the molar mass leads to

G = n Mg (T, p) = H − T S. (24.142)

Thus, the molar specific Gibbs energy is

G
Gm = = Mg (T, p) = Hm − T Sm (24.143)
n

Accordingly, the molar specific Gibbs energy for any substance can be calculated
following Tables B.1–B.29 in Appendix B.
24.3 Gibbs Energy 823

24.3.3 Motivation

Closed System

Imagine a closed system, in which an isobaric/isothermal process takes place.


According to Fig. 24.8, the question is, how much non-volume work δWnon-v can
be released by such a system. However, this non-volume work covers any mechan-
ical, electrical or chemical related work. Thus, also the dissipation. The first law of
thermodynamics obeys11
δ Q + δW = dU. (24.144)

The partial energy equation reads as

δW = δWnon-v + δWV = δWnon-v − p dV. (24.145)

Applying the second law of thermodynamics

δ Q + δ
= T dS (24.146)

finally brings the released specific work δWnon-v

δWnon-v = dU + p dV − T dS + δ
. (24.147)

A combination with the definition of the Gibbs energy, see Eq. 24.136, results in

dG = dU + V d p + p dV − T dS − S dT = δWnon-v − δ
+ V d p − S dT.
(24.148)
With dT = 0 and d p = 0 for an isothermal/isobaric process it is

δWnon-v = dG + δ
(24.149)

In case the change of state is reversible, the Gibbs energy is the maximum non-volume
work, i.e.
δWnon-v,rev = dG (24.150)

However, for a single component fluid at isobaric/isothermal conditions it is, cf. Sect.
24.3.1,
δWnon-v = δ
(24.151)

since dG = m dg = m · (−s dT + v d p) = 0. A (reactive) multi-component system


is required, see Sects. 24.4.1 and 24.4.2, to gain non-volume work.

11 Kinetic/potential energies ignored.


824 24 Chemical Reactions

Fig. 24.9 Gibbs


energy—open system

Open System

Imagine an open system, in which an isobaric/isothermal process takes place.


According to Fig. 24.9, the question is, how much non-pressure work δWnon-p can
be released by such a system. However, this non-pressure work covers any mechan-
ical, electrical or chemical related work. Thus, also the dissipation. The first law of
thermodynamics obeys12
δ Q + δWt = dH. (24.152)

The partial energy equation reads as

δWt = δWnon-p + δY = δWnon-p + V d p (24.153)

i.e. the technical work is the sum of pressure work and non-pressure work. Applying
the second law of thermodynamics

δ Q + δ
= T dS (24.154)

finally brings the released specific work δWnon-p

δWnon-p = dH − V d p − T dS + δ
. (24.155)

A combination with the definition of the Gibbs energy, cf. Eq. 24.136, results in

dG = dH − T dS − S dT = δWnon-p − δ
+ V d p − S dT. (24.156)

With dT = 0 and d p = 0 for an isothermal/isobaric process it is

δWnon-p = dG + δ
. (24.157)

In case the change of state is reversible, the Gibbs energy is the maximum non-
pressure work, i.e.
δWnon-p,rev = dG (24.158)

12 Kinetic/potential energies ignored.


24.3 Gibbs Energy 825

In other words
Wnon-p,rev = G 2 − G 1 (24.159)

Thus, in order to calculate the maximum power, that can be released according to
Fig. 24.9, the isobaric/isothermal reaction needs to be reversible, i.e. δ
= 0:
 
Pnon-p,rev = ṅ j G m, j − ṅ i G m,i
j i
  (24.160)
= ṅ j Hm, j − T Sm, j − ṅ i Hm,i − T Sm,i .
j i

However, for a single component fluid at isobaric/isothermal conditions it is, cf. Sect.
24.3.1,
δWnon-p = δ
(24.161)

since dG = m dg = 0. A (reactive) multi-component system is required, see Sects.


24.4.1 and 24.4.2, to gain non-pressure work.

24.4 Chemical Potential

24.4.1 Multi-Component Systems

Now, the focus is on a system in which there are k different substances. Let n 1 be the
number of moles of substance 1, n 2 of substance 2 and so on. For a single-component
fluid, the extensive Gibbs energy G depends only on T and p, see Sect. 24.3. For
variable compositions, however, the extensive Gibbs energy then is

G = G (T, p, n 1 , n 2 , . . . , n k ) . (24.162)

Thus, the total differential follows


    k 
 
∂G ∂G ∂G
dG = dT + dp + dn i . (24.163)
∂T p,n i ∂p T,n i i=1
∂n i T, p,n j

This expression can be interpreted as follows13 :


• Temperature potential
 
∂G
dT → thermal potential (24.164)
∂T p,n i

13For isobaric/isothermal changes of state, the Gibbs free energy can thus only be obtained from a
chemical potential.
826 24 Chemical Reactions

• Pressure potential
 
∂G
d p → mechanical potential (24.165)
∂p T,n i

• Chemical potential

k 
 
∂G
dn i → driver for change in composition (24.166)
i=1
∂n i T, p,n j

This expression is only relevant if the chemical composition of a system respec-


tively the substances vary, i.e. dn i = 0. A single component fluid is characterised
by dn i = 0. Thus, a chemical potential μi is defined by
 
∂G
μi = (24.167)
∂n i T, p,n j

As known from a single component system with dn i = 0, cf. Sect. 24.3.1, the first
two partial differentials in Eq. 24.163 for constant molar quantity follow
 
∂G
= −S (24.168)
∂T p,n i

and  
∂G
= V. (24.169)
∂p T,n i

Hence, it is

k
dG = V d p − S dT + μi dn i . (24.170)
i=1

The chemical potential μi is needed for systems in which the composition changes.
As already known, temperature is the driver for a thermal balancing process, whereas
pressure is the driver for a mechanical balancing process. Consequently, the chemical
potential is the driver for a chemical balancing process.
For any extensive state variable Z in a mixture under isothermal/isobaric condi-
tions it is
 
Z (T, p, xi ) = n xi Z m,i (T, p) = n i Z m,i (T, p) =
i i (24.171)
= n 1 Z m,1 (T, p) + n 2 Z m,2 (T, p) + · · · + n k Z m,k (T, p)
24.4 Chemical Potential 827

with xi being the mole fraction of each component and Z m,i the molar state value.
This can be easily illustrated with the following example

V (T, p) = n 1 Vm,1 (T, p) + n 2 Vm,2 (T, p) + · · · (24.172)

respectively

H (T, p) = n 1 Hm,1 (T, p) + n 2 Hm,2 (T, p) + · · · (24.173)

The entropy must be calculated with care to take into account the irreversibility of the
mixing process. As shown in Chap. 19, one has to calculate with the partial pressure
pi of each component in the mixture, i.e.

S (T, p) = n 1 Sm,1 (T, p1 ) + n 2 Sm,2 (T, p2 ) + · · · (24.174)

For isothermal/isobaric conditions it results in


 
∂Z
= Z m,i . (24.175)
∂n i p,T,n j

Thus, substituting Z with G and the molar variable Z m,i with G m,i leads to
 
∂G
μi = = G m,i = Hm,i − T Sm,i . (24.176)
∂n i T, p,n j

The temperature dependency of μi is


 
∂μi ∂2G ∂ ∂G ∂
= = =− S = −Sm,i . (24.177)
∂T ∂ T ∂n i ∂n i ∂ T ∂n i

The pressure dependency of μi is

∂μi ∂2G ∂ ∂G ∂
= = = V = Vm,i . (24.178)
∂p ∂ p∂n i ∂n i ∂ p ∂n i

Hence, if T = const. it follows

dμi = Vm,i d p. (24.179)

Based on the ideal gas equation pi V = n i RM T and a mixture of ideal gases, i.e.
pi
p
= VVi , it is
pi n i RM T RM T
Vi = V = ⇒ Vm,i = . (24.180)
p p p

The integration from standard state (0) to partial pressure pi yields


828 24 Chemical Reactions

pi
μi (T ) = μi0 (T ) + RM T ln (24.181)
p0

Note that the standard state is different at different temperatures. As already men-
tioned, it is  
∂G
μi = = G m,i = Hm,i − T Sm,i . (24.182)
∂n i T, p,n j

Finally, the Gibbs-enthalpy for isothermal/isobaric conditions results in


  
G= n i G m,i = n i μi (T, p, xi ) = n i Hm,i − T Sm,i (24.183)
i i i

The same result occurs if


k
dG = V d p − S dT + μi dn i (24.184)
i=1

is integrated while p and T remain constant.

Fundamental Equations

It has been shown that


k
dG = V d p − S dT + μi dn i (24.185)
i=1

However, it is by definition
G = H − T S. (24.186)

Thus, the differential reads as

dG = dH − T dS − S dT. (24.187)

This yields the differential for the enthalpy

dH = dG + T dS + S dT. (24.188)

In combination with Eq. 24.185 it results in


24.4 Chemical Potential 829


k
dH = V d p + T dS + μi dn i (24.189)
i=1

Since
H = U + pV (24.190)

it is
dU = dH − p dV − V d p. (24.191)

A combination with Eq. 24.189 obeys


k
dU = − p dV + T dS + μi dn i (24.192)
i=1

In addition, the following total differentials exist:


    k 
 
∂G ∂G ∂G
dG (T, p, n i ) = dT + dp + dn i .
∂T p,n i ∂p T,n i i=1
∂n i T, p,n j
(24.193)
    k 
 
∂H ∂H ∂H
dH ( p, S, n i ) = dp + dS + dn i .
∂p S,n i ∂S p,n i i=1
∂n i p,S,n j
(24.194)
    k 
 
∂U ∂U ∂U
dU (V, S, n i ) = dV + dS + dn i .
∂V S,n i ∂S V,n i i=1
∂n i V,S,n j
(24.195)
Hence a comparison of Eqs. 24.185, 24.189, 24.192 with Eqs. 24.193–24.195 results
in      
∂G ∂H ∂U
μi = = = (24.196)
∂n i T, p,n j ∂n i p,S,n j ∂n i V,S,n j

It also follows  
∂G
= −S (24.197)
∂T p,n i

 
∂G
=V (24.198)
∂p T,n i

 
∂H
=V (24.199)
∂p S,n i
830 24 Chemical Reactions
 
∂H
=T (24.200)
∂S p,n i

 
∂U
= −p (24.201)
∂V S,n i

 
∂U
= T. (24.202)
∂S V,n i

Example 24.4 Consider the mixing of n A moles A with n B moles of B both at the
same temperature T and the same pressure p, cf. Fig. 24.10. According to Dalton it is

pA + pB = p. (24.203)

The first law of thermodynamic obeys

W12 + Q 12 = 0 = U2 − U1
   
= n A Um,A (T2 ) − Um,A (T1 ) + n B Um,B (T2 ) − Um,B (T1 ) .
(24.204)
Thus, it is T2 = T1 = T . Consequently, it follows
   
H = n A Hm,A (T2 ) − Hm,A (T1 ) + n B Hm,B (T2 ) − Hm,B (T1 ) = 0 (24.205)

respectively
H
Hm = = 0. (24.206)
n

Fig. 24.10 Mixing of two ideal gases


24.4 Chemical Potential 831

According to the second law of thermodynamics it is

S = Si,12 + Sa,12 = Si,12 > 0. (24.207)

The increase in entropy is due to the generation of entropy in the transient balancing
process. Since the system is supposed to be adiabatic, no entropy is exchanged with
the environment. The entropy for the mixture therefore obeys
   
S = n A Sm,A (T2 , pA ) − Sm,A (T1 , p) + n B Sm,B (T2 , pB ) − Sm,B (T1 , p) > 0
(24.208)
respectively
S
Sm = > 0. (24.209)
n
The Gibbs energy for the mixing reads as

G = G 2 − G 1 (24.210)

with    
G 1 = n A Hm,A,1 − T Sm,A,1 + n B Hm,B,1 − T Sm,B,1 (24.211)

and    
G 2 = n A Hm,A,2 − T Sm,A,2 + n B Hm,B,2 − T Sm,B,2 . (24.212)

Since T = const. it follows Hm,A,1 = Hm,A,2 and Hm,B,1 = Hm,B,2 . Thus, it results in
   
G = −n A Sm,A,2 − Sm,A,1 T − n B Sm,B,2 − Sm,B,1 T (24.213)

with pA
Sm,A,1 = Sm,A
0
− RM ln (24.214)
p0

and p
Sm,A,2 = Sm,A
0
− RM ln . (24.215)
p0

B follows analogously, i.e.


pA pB
G = n A RM T ln + n B RM T ln (24.216)
p p

Alternatively, a second approach can be applied: The change of Gibbs energy is

G = G 2 − G 1 . (24.217)

With the following Gibbs energies:


832 24 Chemical Reactions

G 2 = n A μA,2 + n B μB,2 (24.218)

and
G 1 = n A μA,1 + n B μB,1 . (24.219)

The chemical potentials are as follows. For gas A before mixing it yields
p
μA,1 = μ0A + RM T ln (24.220)
p0

and after mixing it is


pA
μA,2 = μ0A + RM T ln . (24.221)
p0

For gas B it follows accordingly


p
μB,1 = μ0B + RM T ln (24.222)
p0

and pB
μB,2 = μ0B + RM T ln . (24.223)
p0

Hence, the change of the Gibbs energy results in


pA pB
G = n A RM T ln + n B RM T ln . (24.224)
p p

Since ni pi
= xi = πi = (24.225)
n p

it finally is
G = n RM T (xA lnxA + xB lnxB ) (24.226)

As indicated by Fig. 24.10 the change of Gibbs energy G is negative, so that G


decreases when gases are mixed. A negative G indicates, that the process runs
spontaneously without impact from outside. In case the molar quantity of A is the
same as of B, entropy generation is maximised and the Gibbs energy minimised. The
generated entropy follows Eq. 24.207, i.e.
   
Si,12 = n A Sm,A (T2 , pA ) − Sm,A (T1 , p) + n B Sm,B (T2 , pB ) − Sm,B (T1 , p) > 0.
(24.227)
Thus, the dissipation for the isothermal change of state is


12 = T · Si,12 = −n RM T (xA lnxA + xB lnxB ) . (24.228)
24.4 Chemical Potential 833

However, according to Eq. 24.149 no work Wnon-v,12 can be gained, since the entire
Gibbs energy is consumed by the dissipation:

Wnon-v,12 = G 12 +
12 = 0 (24.229)

Once the mixing is complete, i.e. a thermodynamic equilibrium is reached in state


(2), G is zero, i.e. the system is not underlying a further change of state.

Example 24.5 This example explains an evaporation process, see Sect. 20.1.3,
applying the chemical potential. It has already been shown that the chemical poten-
tial is the driver for chemical reactions or diffusive transport processes. Evaporation
from a water surface is a non-reactive process, cf. Fig. 24.11. Water molecules with
sufficient molecular kinetic energy can detach from the liquid surface and pass into
the gas phase, i.e. air. The driver for such a mass transfer is the chemical potential,
which can be calculated according to Eq. 24.176:
• Chemical potential at (1) for water in the liquid phase

μH2 O,liq = G m,H2 O,liq = Hm,H2 O,liq − T Sm,H2 O,liq (24.230)

• Chemical potential at (1) for water in the vapour phase

μH2 O,v = G m,H2 O,v = Hm,H2 O,v − T Sm,H2 O,v (24.231)

Molar enthalpies as well as entropies are listed in Table B. Both the gaseous and liquid
phases are in thermal, i.e. T is the same for all phases, and mechanical equilibrium.
The pressure in the gaseous phase composed of the partial pressures, i.e.

Fig. 24.11 Chemical potential at evaporation


834 24 Chemical Reactions

ptotal = pa + pv . (24.232)

The pressure in the liquid at the interface is due to mechanical equilibrium ptotal .
Equations (24.230) and (24.231) are plotted in Fig. 24.11 as a function of the vapour
pressure in the gaseous phase. For this example, a temperature of 25 ◦ C and a pres-
sure14 of ptotal = 1 bar was chosen. The chemical potential of liquid water is not
affected by the partial pressure, as its pressure is assumed to be constant. However,
to calculate the chemical potential of the vapour according to Table B, the partial
pressure of the vapour must be considered in the molar entropies.
At the beginning, the air above the liquid is dry, i.e. pv = 0. It can be seen that the
largest chemical potential difference μH2 O exists at this point. The mass transfer
from liquid to gas phase is therefore most intensive. The more water molecules
finally evaporate, the greater the partial pressure of the water vapour in the gas phase.
The driving difference of the chemical potentials becomes smaller until finally the
chemical potential of the vaporous water corresponds to the potential of the liquid
water. This is achieved at saturation pressure ps (T ). This pressure corresponds to the
saturation pressure of water at 25 ◦ C, i.e. 0.0317 bar. At a greater distance from the
surface, e.g. point (2), the partial pressure of the vapour is lower than the saturation
state. The same applies to the chemical potential, i.e.

μH2 O,v,2 < μH2 O,v,1 (24.233)

so that water continues to evaporate from the surface until finally the gas phase is
completely saturated.

24.4.2 Chemical Reactions

Let there be a dynamic chemical reaction under isothermal/isobaric conditions


according to
γ A A + γB B  γC C + γD D (24.234)

The reaction can go back and forth. When a dynamic equilibrium is reached, the
reactions run at the same velocity in both directions.15 However, the Gibbs energy
G reaches a minimum, so that
dG = 0. (24.235)

Macroscopically, the molar quantities of all components remain constant, although


the chemical reactions proceed in both directions. In states (1) and (2), the molar
quantities of the individual components are measured, see Fig. 24.12. The following
cases can occur:

14 Simplifying, the total pressure is considered constant.


15 Hence, it is no static equilibrium. A static equilibrium requires dn i = 0.
24.4 Chemical Potential 835

Fig. 24.12 Dynamic chemical reaction

• If G = G 2 − G 1 < 0, the process runs spontaneously. The formation of products


is dominating.
• If dG = G 2 − G 1 = 0, the process has already reached a chemical equilibrium,
i.e. the Gibbs energy reaches a minimum, see also Example 24.6.
• If G = G 2 − G 1 > 0, the process does not run spontaneously. Energy is required.
The formation of educts is dominating.

Hence, in thermodynamic equilibrium it is


    k 
 
∂G ∂G ∂G
dG = dT + dp + dn i = 0. (24.236)
∂T p,n i ∂p T,n i ∂n i T, p,n j
i=1
 
μi

Since the reaction shall be isothermal/isobaric the equation simplifies to


k
μi dn i = 0. (24.237)
i=1

Applied to the given example, this means

μA dn A + μB dn B + μC dn C + μD dn D = 0. (24.238)

According to the chemical equation it is:


γA γA
nA = − n D → dn A = − dn D (24.239)
γD γD
γB γB
nB = − n D → dn B = − dn D (24.240)
γD γD
γC γC
nC = n D → dn C = dn D (24.241)
γD γD
836 24 Chemical Reactions

Thus, it follows
 
γA γB γC γD
−μA − μB + μC + μD dn D = 0 (24.242)
γD γD γD γD

respectively
− γA μA − γB μB + γC μC + γD μD = 0. (24.243)

In other words, it is
 
γi μi − γjμj = 0 (24.244)
i,products j,educts

Example 24.6 This example shows an equilibrium reaction

N2 (g) + 3H2 (g)  2NH3 (g) (24.245)

An equilibrium reaction can run back and forth, see Fig. 24.13: Educts on the
left hand side are chemically reactive and form the product on the right hand side.
However, the mechanism can run vice versa, so that the product is split into the educts
again. The driver for this reaction is the change of Gibbs energy from products to
educts. In equilibrium state dG = 0 must be satisfied. Thus, it obeys

− μH2 − 3μH2 + 2μNH3 = 0. (24.246)

In an isothermal/isobaric reaction, the cause of the chemical reaction and its equi-
librium state is the chemical potential μ of the components involved.

Fig. 24.13 Chemical


balancing driven by μi
24.5 Exergy of a Fossil Fuel 837

24.5 Exergy of a Fossil Fuel

Let us have a closer look at a combustion process of a fossil fuel. For the following
investigations any changes of kinetic and potential energies shall be ignored. The
combustion shall be performed isothermally at ambient temperature Tenv = 298.15 K
and isobarically at ambient pressure penv = 1 bar, cf. Fig. 24.14. This represents a
standard atmosphere. Since not only the thermomechanical properties of this standard
atmosphere are relevant, its standard chemical composition is also defined according
to [44], see Table 24.1. This standard atmosphere is in thermodynamic equilibrium,
i.e. it is saturated with water. In this saturated atmosphere with a pressure of penv =
1 bar, liquid water occurs further. The combustion shall be carried out with oxygen
extracted from this atmosphere but supplied to the combustion with penv , cf Fig.
24.14. The exhaust gas (PR) is released to the huge atmosphere without changing its
composition significantly. In order to find the exergy of the fossil fuel, the process
must be reversible, see Sect. 24.3.3.
The following balances can be performed for steady state:
• First law of thermodynamics, see Fig. 24.14

Fig. 24.14 Exergy of a fossil fuel—energy balance

Table 24.1 Chemical composition of a standard atmosphere at 25 ◦ C and 1 bar


xenv,N2 xenv,O2 xenv,H2 O xenv,Ar xenv,CO2
0.75608 0.20284 0.03171 0.00906 0.00031

Fig. 24.15 Exergy of a fossil fuel—entropy balance


838 24 Chemical Reactions

Q̇ + Prev + ṅ O2 ,min Hm,O2 + ṅ F Hm,F = ṅ Pr, j Hm, j (24.247)
j

Rearranging brings

Q̇ + Prev = ṅ Pr, j Hm, j − ṅ O2 ,min Hm,O2 − ṅ F Hm,F . (24.248)
j

If the product water is in the liquid state, the heat flux and the released power are
equal to the specific higher heating value, i.e.

Q̇ + Prev = −ṅ F H0M (Tenv ) . (24.249)

• Second law of thermodynamics, see Fig. 24.15



Ṡi + Ṡa + ṅ O2 ,min Sm,O2 + ṅ F Sm,F = ṅ Pr, j Sm, j (24.250)
j

Since the technical power released is to be maximised, the reaction must be


reversible so that
Ṡi = 0. (24.251)

The exchanged entropy with the heat Ṡa can easily be substituted, since the com-
bustion is isothermal, i.e.

Ṡa = . (24.252)
Tenv

Hence, the balance of entropy is

Q̇ 
= ṅ Pr, j Sm, j − ṅ O2 ,min Sm,O2 − ṅ F Sm,F . (24.253)
Tenv j

The heat flux then is


⎡ ⎤

Q̇ = Tenv ⎣ ṅ Pr, j Sm, j − ṅ O2 ,min Sm,O2 − ṅ F Sm,F ⎦ . (24.254)
j

In combination with the first law of thermodynamics, cf. Eq. 24.248, it results in

Prev = ṅ Pr, j Hm, j − ṅ O2 ,min Hm,O2 − ṅ F Hm,F +
j
⎡ ⎤
 (24.255)
− Tenv ⎣ ṅ Pr, j Sm, j − ṅ O2 ,min Sm,O2 − ṅ F Sm,F ⎦ .
j
24.5 Exergy of a Fossil Fuel 839

With the newly introduced Gibbs energy it is



Prev = ṅ Pr, j G m, j − ṅ O2 ,min G m,O2 − ṅ F G m,F (24.256)
j

respectively
Prev 
= νPr, j G m, j − Omin G m,O2 − G m,F . (24.257)
ṅ F j

• Exergy balance, see Fig. 24.16



Ė x,Q + Prev + ṅ F E xm,F + ṅ O2 ,min E xm,O2 =  Ė x,V + ṅ Pr, j E xm, j (24.258)
j

Since the heat Q̇ is transferred at environmental temperature Tenv it is

Ė x,Q = 0. (24.259)

The process is further supposed to be reversible so that  Ė x,V = 0. Solving for


the exergy E xm,F of the fuel leads to

ṅ F E xm,F = −Prev + ṅ Pr, j E xm, j − ṅ O2 ,min E xm,O2 (24.260)
j

respectively

Prev 
E xm,F = − + νPr, j E xm, j − Omin E xm,O2 . (24.261)
ṅ F j

Fig. 24.16 Exergy of a fossil fuel—exergy balance


840 24 Chemical Reactions

The exergy of the flows has been treated in Sect. 16.2. A flow in Part I used to
carry exergy due to a thermal and mechanical16 imbalance with the environment.
The molar specific exergy of oxygen is therefore given by
 
E xm,O2 = Hm,O2 − Hm,O2 ,env +Tenv Sm,O2 ,env − Sm,O2 . (24.262)
 
=0

The first term disappears, since the oxygen enters with ambient temperature17 and
enthalpy purely depends on temperature. The oxygen is supplied to the combustion
with a pressure of penv , see Fig. 24.16:

pO2 = penv . (24.263)

It therefore contains exergy because the partial pressure of oxygen is lower in the
standard atmosphere:
penv,O2 = penv xenv,O2 . (24.264)

The exergy of the flow of oxygen obeys18


 
Tenv penv xenv,O2 Tenv penv
E xm,O2 = 0 + Tenv C p ln − RM ln − C p ln + RM ln
T0 p0 T0 p0
(24.265)
respectively
 
penv 1
E xm,O2 = 0 + Tenv RM ln = Tenv RM ln . (24.266)
penv xenv,O2 xenv,O2

The oxygen at the inlet therefore has exergy due to its chemical potential compared
to the standard atmosphere.

The products can be treated accordingly: However, during the combustion the
exhaust gas components would mix with one another and leave the combustion
process as a mixture of gases with a total pressure of penv . Each component in this
mixture then has a partial pressure that correlates with its concentration. Unfortu-
nately, such a mixture is related with generation of entropy and thus with a loss
of exergy, see Sect. 19.3.4. In order to find the exergy, i.e. maximum working
capability of the fuel, any irreversibility has to be avoided. Consequently, in this
approach each product component leaves the system boundary with ambient tem-
perature and ambient pressure penv , i.e. no mixing occurs, see also Chap. 19 and

16 In case kinetic/potential energies are counted as well.


17 Oxygen in the standard atmosphere and the (pure) oxygen supplied have the same temperature,
see Fig. 24.16.
18 The temperature dependence disappears since supplied oxygen and standard atmosphere have

the same temperature.


24.5 Exergy of a Fossil Fuel 841

Fig. 24.16:
p j = penv . (24.267)

The exergy carried by each component is thus due to the different partial pressure
of this component in the standard atmosphere, i.e. due to a thermomechanical
imbalance. The individual partial pressure of a component j in the environment is

penv, j = penv xenv, j . (24.268)

The exergy then obeys


 
E xm, j = Hm, j − Hm, j,env + Tenv Sm, j,env − Sm, j
 
penv xenv, j
= 0 + Tenv −RM ln
penv (24.269)
1
= Tenv RM ln .
xenv, j

The following then applies to the molar specific exergy of the fuel
⎡ ⎤
Prev  1 1 ⎦
E xm,F = − + Tenv RM ⎣ νPr, j ln − Omin ln (24.270)
ṅ F j
xenv, j xenv,O2

In case the fuel contains sulphur, so that SO2 is produced, the according reference
atmosphere needs to be extended to xenv,SO2 .

Problem 24.3 Calculate the molar specific exergy of methane at standard condi-
tions. I.e. Tenv = 298.15 K, penv = 1 bar and a composition according to Table 24.1.
The process shall be isobaric/isothermal.

Solution

The chemical equation reads as

CH4 + 2 O2 → CO2 + 2 H2 O (24.271)

According to the method introduced in Sect. 23.3.5, the following results:


• Minimum oxygen need:
Omin = 2 (24.272)

• Product composition:
νPr,CO2 = 1 (24.273)
842 24 Chemical Reactions

and
νPr,H2 O = 2 (24.274)

The molar concentrations then follow with νPr = 3:


νPr,CO2
xPr,CO2 = = 0.333 (24.275)
νPr

and νPr,H2 O
xPr,H2 O = = 0.667. (24.276)
νPr
ps (Tenv )
Thus, almost the entire product water is in liquid state, since xPr,H2 O > penv
.
The molar specific Gibbs energies are, refer to Tables B.1–B.29 in Appendix B:
• Methane

kJ
G m,CH4 = Hm,CH4 − Tenv Sm,CH4 = −1.3017 × 105 (24.277)
kmol
• Oxygen

kJ
G m,O2 = Hm,O2 − Tenv Sm,O2 = −6.1165 × 104 (24.278)
kmol
• Carbon dioxide

kJ
G m,CO2 = Hm,CO2 − Tenv Sm,CO2 = −4.5725 × 105 (24.279)
kmol
• Water (liquid)

kJ
G m,H2 O = Hm,H2 O − Tenv Sm,H2 O = −3.0668 × 105 (24.280)
kmol
Thus, it is

Prev  kJ
= νPr, j G m, j − Omin G m,O2 − G m,F = −8.1812 × 105 . (24.281)
ṅ F j
kmol

Finally, the molar specific exergy of the fuel is


⎡ ⎤
Prev  1 1
E xm,F =− + Tenv RM ⎣ νPr, j ln − Omin ln ⎦. (24.282)
ṅ F j
x env, j x env,O2

In this case, the product water is assumed to be completely liquid, it is already in


equilibrium with the liquid water in the atmosphere, which also has a pressure of
24.5 Exergy of a Fossil Fuel 843

penv = 1 bar, i.e.


 
Prev 1 1
E xm,F =− + Tenv RM νPr,CO2 ln − Omin ln
ṅ F xenv,CO2 xenv,O2
(24.283)
kJ
= 830240 .
kmol
Note that the molar specific higher heating value is
 kJ
H0M (Tenv ) = − νPr, j Hm, j + Omin Hm,O2 + Hm,F = 890570 (24.284)
j
kmol

and the molar specific lower heating value with the water being entirely in gaseous
state
 kJ
HUM (Tenv ) = − νPr, j Hm, j + Omin Hm,O2 + Hm,F = 802562 . (24.285)
j
kmol
Appendix A
Steam Table (Water) According to IAPWS

See1 Tables A.1, A.2, A.3, A.4, A.5, A.6, A.7, A.8, A.9, A.10, A.11 and A.12.
(continued)

1International Association for the Properties of Water and Steam: The listed data has been generated
with XSteam, see [9].
© Springer Nature Switzerland AG 2022 845
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
Table A.1 Saturated liquid and saturated steam 1/4
846

ϑ ps h h  s s  v v  cp cp
◦C bar kJ kJ m3 kJ
kg kg K kg kg K

0 0.0061122 −0.041588 2500.9 −0.000155 9.1558 0.0010 206.1431 4.22 1.888


0.01 0.0061166 0.00061178 2500.911 0 9.1555 0.0010002 205.9975 4.2199 1.8882
1 0.0065709 4.1767 2502.7299 0.01526 9.1291 0.0010001 192.4447 4.2165 1.8889
2 0.0070599 8.3916 2504.5666 0.030606 9.1027 0.0010001 179.7636 4.2134 1.8895
3 0.0075808 12.6035 2506.4024 0.045886 9.0765 0.0010001 168.0141 4.2105 1.8902
4 0.0081355 16.8127 2508.2375 0.061101 9.0506 0.0010001 157.1213 4.2078 1.8909
5 0.0087257 21.0194 2510.0717 0.076252 9.0249 0.0010001 147.0169 4.2054 1.8917
6 0.0093535 25.2237 2511.9051 0.09134 8.9994 0.0010001 137.6382 4.2031 1.8924
7 0.010021 29.4258 2513.7377 0.10637 8.9742 0.0010001 128.9281 4.2011 1.8932
8 0.01073 33.626 2515.5693 0.12133 8.9492 0.0010002 120.8344 4.1992 1.894
9 0.011483 37.8244 2517.4001 0.13624 8.9244 0.0010003 113.3092 4.1974 1.8949
10 0.012282 42.0211 2519.2298 0.15109 8.8998 0.0010003 106.3087 4.1958 1.8957
11 0.013129 46.2162 2521.0586 0.16587 8.8755 0.0010004 99.7927 4.1943 1.8966
12 0.014028 50.41 2522.8864 0.18061 8.8514 0.0010005 93.7243 4.193 1.8975
13 0.014981 54.6024 2524.7132 0.19528 8.8275 0.0010007 88.0698 4.1917 1.8985
14 0.015989 58.7936 2526.5389 0.2099 8.8038 0.0010008 82.7981 4.1905 1.8994
15 0.017057 62.9837 2528.3636 0.22447 8.7804 0.0010009 77.8807 4.1894 1.9004
16 0.018188 67.1727 2530.1871 0.23898 8.7571 0.0010011 73.2915 4.1884 1.9014
17 0.019383 71.3608 2532.0094 0.25344 8.7341 0.0010013 69.0063 4.1875 1.9025
18 0.020647 75.5479 2533.8307 0.26785 8.7112 0.0010015 65.0029 4.1866 1.9035
19 0.021982 79.7343 2535.6507 0.2822 8.6886 0.0010016 61.2609 4.1858 1.9046
20 0.023392 83.9199 2537.4695 0.2965 8.6661 0.0010018 57.7615 4.1851 1.9057
21 0.024881 88.1048 2539.287 0.31075 8.6439 0.0010021 54.4873 4.1844 1.9069
22 0.026452 92.289 2541.1033 0.32495 8.6218 0.0010023 51.4225 4.1838 1.908
Appendix A: Steam Table (Water) According to IAPWS
Table A.2 Saturated liquid and saturated steam 2/4
ϑ ps h h  s s  v v  cp cp
◦C bar kJ kJ m3 kJ
kg kg K kg kg K
23 0.028109 96.4727 2542.9182 0.3391 8.6 0.0010025 48.5521 4.1832 1.9092
24 0.029856 100.6558 2544.7318 0.3532 8.5783 0.0010028 45.8626 4.1827 1.9104
25 0.031697 104.8384 2546.5441 0.36726 8.5568 0.001003 43.3414 4.1822 1.9116
26 0.033637 109.0205 2548.3549 0.38126 8.5355 0.0010033 40.9768 4.1817 1.9129
27 0.035679 113.2022 2550.1643 0.39521 8.5144 0.0010035 38.7582 4.1813 1.9141
28 0.037828 117.3835 2551.9723 0.40912 8.4934 0.0010038 36.6754 4.1809 1.9154
29 0.040089 121.5645 2553.7788 0.42298 8.4727 0.0010041 34.7194 4.1806 1.9167
30 0.042467 125.7452 2555.5837 0.43679 8.4521 0.0010044 32.8816 4.1803 1.918
31 0.044966 129.9255 2557.3871 0.45056 8.4317 0.0010047 31.154 4.18 1.9194
32 0.047592 134.1057 2559.189 0.46428 8.4115 0.001005 29.5295 4.1798 1.9207
33 0.050351 138.2855 2560.9892 0.47795 8.3914 0.0010054 28.001 4.1795 1.9221
34 0.053247 142.4653 2562.7877 0.49158 8.3715 0.0010057 26.5624 4.1794 1.9235
Appendix A: Steam Table (Water) According to IAPWS

35 0.056286 146.6448 2564.5846 0.50517 8.3518 0.001006 25.2078 4.1792 1.9249


36 0.059475 150.8242 2566.3797 0.51871 8.3323 0.0010064 23.9318 4.1791 1.9263
37 0.062818 155.0035 2568.1731 0.5322 8.3129 0.0010068 22.7292 4.179 1.9278
38 0.066324 159.1827 2569.9647 0.54566 8.2936 0.0010071 21.5954 4.1789 1.9292
39 0.069997 163.3619 2571.7545 0.55906 8.2746 0.0010075 20.5261 4.1788 1.9307
40 0.073844 167.541 2573.5424 0.57243 8.2557 0.0010079 19.517 4.1788 1.9322
41 0.077873 171.7202 2575.3284 0.58575 8.2369 0.0010083 18.5646 4.1788 1.9337
42 0.08209 175.8993 2577.1125 0.59903 8.2183 0.0010087 17.6652 4.1788 1.9353
43 0.086503 180.0785 2578.8946 0.61227 8.1999 0.0010091 16.8155 4.1788 1.9368
44 0.091118 184.2578 2580.6746 0.62547 8.1816 0.0010095 16.0126 4.1789 1.9384
45 0.095944 188.4372 2582.4526 0.63862 8.1634 0.0010099 15.2534 4.179 1.94
847
Table A.2 (continued)
848

ϑ ps h h  s s  v v  cp cp
◦C bar kJ kJ m3 kJ
kg kgK kg kgK
46 0.10099 192.6167 2584.2285 0.65174 8.1454 0.0010103 14.5355 4.1791 1.9416
47 0.10626 196.7963 2586.0023 0.66481 8.1276 0.0010108 13.8562 4.1792 1.9432
48 0.11176 200.9761 2587.7739 0.67785 8.1099 0.0010112 13.2132 4.1794 1.9449
49 0.11751 205.156 2589.5432 0.69084 8.0923 0.0010117 12.6045 4.1796 1.9466
50 0.12351 209.3362 2591.3103 0.70379 8.0749 0.0010121 12.0279 4.1798 1.9482
51 0.12977 213.5166 2593.075 0.71671 8.0576 0.0010126 11.4815 4.18 1.95
52 0.13631 217.6973 2594.8374 0.72958 8.0405 0.0010131 10.9637 4.1802 1.9517
53 0.14312 221.8782 2596.5973 0.74242 8.0235 0.0010136 10.4726 4.1805 1.9534
54 0.15022 226.0594 2598.3548 0.75522 8.0066 0.001014 10.0069 4.1808 1.9552
55 0.15761 230.241 2600.1098 0.76798 7.9899 0.0010145 9.5649 4.1811 1.957
56 0.16532 234.4229 2601.8622 0.7807 7.9733 0.001015 9.1454 4.1814 1.9588
57 0.17335 238.6052 2603.6121 0.79339 7.9568 0.0010155 8.7471 4.1818 1.9607
58 0.18171 242.7878 2605.3592 0.80603 7.9405 0.0010161 8.3688 4.1821 1.9625
59 0.19041 246.9709 2607.1037 0.81864 7.9243 0.0010166 8.0093 4.1825 1.9644
60 0.19946 251.1544 2608.8454 0.83122 7.9082 0.0010171 7.6677 4.1829 1.9664
61 0.20887 255.3383 2610.5843 0.84375 7.8922 0.0010176 7.3428 4.1834 1.9683
62 0.21866 259.5228 2612.3203 0.85625 7.8764 0.0010182 7.0338 4.1838 1.9703
63 0.22884 263.7077 2614.0534 0.86872 7.8607 0.0010187 6.7399 4.1843 1.9723
64 0.23942 267.8931 2615.7836 0.88115 7.8451 0.0010193 6.4601 4.1848 1.9743
65 0.25041 272.0791 2617.5107 0.89354 7.8296 0.0010199 6.1938 4.1853 1.9764
66 0.26183 276.2657 2619.2347 0.9059 7.8142 0.0010204 5.9402 4.1859 1.9785
67 0.27368 280.4528 2620.9556 0.91823 7.799 0.001021 5.6986 4.1864 1.9806
68 0.28599 284.6405 2622.6733 0.93052 7.7839 0.0010216 5.4684 4.187 1.9828
69 0.29876 288.8289 2624.3877 0.94277 7.7689 0.0010222 5.249 4.1876 1.985
70 0.31201 293.0179 2626.0988 0.95499 7.754 0.0010228 5.0397 4.1882 1.9873
Appendix A: Steam Table (Water) According to IAPWS
Table A.3 Saturated liquid and saturated steam 3/4
ϑ ps h h  s s  v v  cp cp
◦C kJ kJ m3 kJ
bar kg kg K kg kg K
71 0.32575 297.2076 2627.8066 0.96718 7.7392 0.0010234 4.8402 4.1889 1.9895
72 0.34 301.398 2629.5109 0.97933 7.7245 0.001024 4.6498 4.1896 1.9919
73 0.35478 305.5891 2631.2117 0.99146 7.71 0.0010246 4.4681 4.1902 1.9942
74 0.37009 309.781 2632.909 1.0035 7.6955 0.0010252 4.2947 4.191 1.9966
75 0.38595 313.9736 2634.6026 1.0156 7.6812 0.0010258 4.1291 4.1917 1.999
76 0.40239 318.1669 2636.2926 1.0276 7.6669 0.0010265 3.9709 4.1924 2.0015
77 0.41941 322.3611 2637.9788 1.0396 7.6528 0.0010271 3.8198 4.1932 2.0041
78 0.43703 326.5561 2639.6612 1.0516 7.6388 0.0010277 3.6754 4.194 2.0066
79 0.45527 330.752 2641.3397 1.0635 7.6248 0.0010284 3.5373 4.1948 2.0092
80 0.47415 334.9487 2643.0143 1.0754 7.611 0.001029 3.4053 4.1956 2.0119
Appendix A: Steam Table (Water) According to IAPWS

81 0.49368 339.1463 2644.6849 1.0873 7.5973 0.0010297 3.279 4.1965 2.0146


82 0.51387 343.3448 2646.3515 1.0991 7.5837 0.0010304 3.1582 4.1974 2.0174
83 0.53476 347.5443 2648.0139 1.1109 7.5701 0.001031 3.0426 4.1983 2.0202
84 0.55636 351.7447 2649.672 1.1227 7.5567 0.0010317 2.9319 4.1992 2.0231
85 0.57867 355.9461 2651.3259 1.1344 7.5434 0.0010324 2.8259 4.2001 2.026
86 0.60174 360.1485 2652.9755 1.1461 7.5301 0.0010331 2.7244 4.2011 2.029
87 0.62556 364.3519 2654.6206 1.1578 7.517 0.0010338 2.6272 4.202 2.0321
88 0.65017 368.5563 2656.2613 1.1694 7.5039 0.0010345 2.5341 4.203 2.0352
89 0.67559 372.7618 2657.8973 1.1811 7.4909 0.0010352 2.4448 4.2041 2.0383
(continued)
849
850

Table A.3 (continued)


ϑ ps h h  s s  v v  cp cp
◦C kJ kJ m3 kJ
bar kg kg K kg kg K
90 0.70182 376.9684 2659.5288 1.1927 7.4781 0.0010359 2.3591 4.2051 2.0415
91 0.7289 381.1762 2661.1555 1.2042 7.4653 0.0010367 2.2771 4.2062 2.0448
92 0.75685 385.385 2662.7775 1.2158 7.4526 0.0010374 2.1983 4.2072 2.0482
93 0.78568 389.595 2664.3946 1.2273 7.44 0.0010381 2.1228 4.2083 2.0516
94 0.81542 393.8062 2666.0068 1.2387 7.4275 0.0010389 2.0502 4.2095 2.0551
95 0.84609 398.0185 2667.6139 1.2502 7.415 0.0010396 1.9806 4.2106 2.0586
96 0.87771 402.2321 2669.216 1.2616 7.4027 0.0010404 1.9138 4.2118 2.0623
97 0.91031 406.447 2670.813 1.273 7.3904 0.0010411 1.8497 4.213 2.066
98 0.9439 410.6631 2672.4047 1.2844 7.3782 0.0010419 1.788 4.2142 2.0697
99 0.97852 414.8805 2673.991 1.2957 7.3661 0.0010427 1.7288 4.2154 2.0736
100 1.0142 419.0992 2675.572 1.307 7.3541 0.0010435 1.6719 4.2166 2.0775
105 1.209 440.2131 2683.3933 1.3632 7.2951 0.0010474 1.4185 4.2232 2.0983
110 1.4338 461.3634 2691.0676 1.4187 7.238 0.0010516 1.2094 4.2304 2.1212
115 1.6918 482.5528 2698.5848 1.4735 7.1827 0.0010559 1.0359 4.2381 2.1464
120 1.9867 503.7846 2705.9342 1.5278 7.1291 0.0010603 0.8913 4.2464 2.174
125 2.3222 525.0618 2713.1055 1.5815 7.077 0.0010649 0.77011 4.2553 2.2042
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.3 (continued)
ϑ ps h h  s s  v v  cp cp
◦C kJ kJ m3 kJ
bar kg kg K kg kg K
130 2.7026 546.3878 2720.0878 1.6346 7.0264 0.0010697 0.66808 4.2648 2.237
135 3.132 567.7661 2726.8708 1.6872 6.9772 0.0010747 0.5818 4.2751 2.2726
140 3.615 589.2003 2733.4439 1.7393 6.9293 0.0010798 0.50852 4.286 2.3109
145 4.1563 610.6941 2739.7968 1.7909 6.8826 0.001085 0.44602 4.2978 2.352
150 4.761 632.2516 2745.9191 1.842 6.837 0.0010905 0.3925 4.3103 2.3959
155 5.4342 653.8769 2751.8005 1.8926 6.7926 0.0010962 0.3465 4.3236 2.4425
160 6.1814 675.5747 2757.4305 1.9428 6.7491 0.001102 0.30682 4.3379 2.4918
Appendix A: Steam Table (Water) According to IAPWS

165 7.0082 697.3495 2762.7985 1.9926 6.7066 0.001108 0.27246 4.3532 2.5438
170 7.9205 719.2064 2767.8937 2.0419 6.6649 0.0011143 0.24262 4.3695 2.5985
175 8.9245 741.1507 2772.7045 2.0909 6.6241 0.0011207 0.2166 4.3869 2.656
180 10.0263 763.188 2777.2194 2.1395 6.5841 0.0011274 0.19386 4.4056 2.7164
185 11.2327 785.3243 2781.4259 2.1878 6.5447 0.0011343 0.17392 4.4255 2.7798
190 12.5502 807.566 2785.311 2.2358 6.506 0.0011414 0.15638 4.4468 2.8464
851
852

Table A.4 Saturated liquid and saturated steam 4/4


ϑ ps h h  s s  v v  cp cp
◦C kJ kJ m3 kJ
bar kg kg K kg kg K
195 13.9858 829.9199 2788.861 2.2834 6.4679 0.0011488 0.14091 4.4696 2.9163
200 15.5467 852.3931 2792.0616 2.3308 6.4303 0.0011565 0.12722 4.494 2.99
205 17.2402 874.9933 2794.8974 2.3779 6.3932 0.0011645 0.11509 4.5202 3.0677
210 19.0739 897.7289 2797.3523 2.4248 6.3565 0.0011727 0.1043 4.5482 3.1496
215 21.0555 920.6086 2799.4096 2.4714 6.3202 0.0011813 0.094689 4.5784 3.2363
220 23.1929 943.6417 2801.051 2.5178 6.2842 0.0011902 0.086101 4.6107 3.328
225 25.4942 966.8384 2802.2577 2.5641 6.2485 0.0011994 0.078411 4.6455 3.4252
230 27.9679 990.2095 2803.0093 2.6102 6.2131 0.001209 0.07151 4.6829 3.5283
235 30.6224 1013.7668 2803.2844 2.6561 6.1777 0.001219 0.065304 4.7233 3.6379
240 33.4665 1037.5228 2803.06 2.7019 6.1425 0.0012295 0.05971 4.7668 3.7545
245 36.5091 1061.4911 2802.3114 2.7477 6.1074 0.0012404 0.054658 4.8139 3.8789
250 39.7594 1085.6868 2801.0121 2.7934 6.0722 0.0012517 0.050087 4.865 4.0119
255 43.2267 1110.126 2799.1333 2.8391 6.037 0.0012636 0.045941 4.9204 4.1545
260 46.9207 1134.8266 2796.6436 2.8847 6.0017 0.0012761 0.042175 4.9807 4.308
265 50.8512 1159.8083 2793.5086 2.9304 5.9662 0.0012892 0.038748 5.0466 4.4741
270 55.0284 1185.0928 2789.6899 2.9762 5.9304 0.001303 0.035622 5.1188 4.6547
275 59.4626 1210.7046 2785.1446 3.0221 5.8943 0.0013175 0.032767 5.1982 4.8524
280 64.1646 1236.671 2779.8245 3.0681 5.8578 0.0013328 0.030154 5.2859 5.0701
285 69.1454 1263.023 2773.6747 3.1143 5.8208 0.0013491 0.027758 5.3832 5.3114
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.4 (continued)
ϑ ps h h  s s  v v  cp cp
◦C kJ kJ m3 kJ
bar kg kg K kg kg K
290 74.4164 1289.7957 2766.6326 3.1608 5.7832 0.0013663 0.025557 5.4918 5.5806
295 79.9895 1317.0297 2758.6266 3.2076 5.7449 0.0013846 0.023531 5.6137 5.8825
300 85.8771 1344.7713 2749.5737 3.2547 5.7058 0.0014042 0.021663 5.7515 6.2231
305 92.0919 1373.0748 2739.3776 3.3024 5.6656 0.0014252 0.019937 5.9083 6.6096
310 98.6475 1402.0034 2727.9243 3.3506 5.6243 0.0014479 0.018339 6.0883 7.0513
315 105.558 1431.6321 2715.0772 3.3994 5.5816 0.0014724 0.016856 6.2968 7.561
320 112.8386 1462.051 2700.6677 3.4491 5.5373 0.0014991 0.015476 6.5414 8.1575
325 120.5052 1493.3719 2684.483 3.4997 5.4911 0.0015283 0.014189 6.8331 8.8689
330 128.5752 1525.738 2666.2483 3.5516 5.4425 0.0015606 0.012984 7.1888 9.7381
335 137.0673 1559.3407 2645.6023 3.6048 5.391 0.0015967 0.011852 7.6354 10.83
Appendix A: Steam Table (Water) According to IAPWS

340 146.0018 1594.4466 2622.0667 3.6599 5.3359 0.0016375 0.010784 8.2166 12.2412
345 155.4015 1631.4365 2595.0082 3.7175 5.2763 0.0016846 0.0097698 9.0023 14.112
350 165.2916 1670.8892 2563.6305 3.7783 5.2109 0.0017401 0.0088009 10.102 16.6415
355 175.7012 1713.7096 2526.4499 3.8438 5.1377 0.0018078 0.007866 11.8584 20.7136
360 186.664 1761.4911 2480.9862 3.9164 5.0527 0.0018945 0.006945 14.8744 27.5691
365 198.2216 1817.5893 2422.0051 4.001 4.9482 0.0020156 0.0060043 21.4744 42.0135
370 210.4337 1892.6429 2333.5016 4.1142 4.7996 0.0022221 0.0049462 47.1001 93.4065
373.946 220.6397 2087.5 4.412 0.0031 ∞
853
854

m3
Table A.5 Specific volume v of water in kg 1/2
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
0.01 0.0010002 137.5362 149.0961 160.6471 172.1934 183.737 195.2789 218.3599 241.4391 264.5173 287.5949
0.1 0.0010002 0.001003 14.8674 16.0347 17.1967 18.356 19.5136 21.826 24.1365 26.446 28.755
1 0.0010002 0.001003 0.0010121 0.0010258 1.696 1.8173 1.9367 2.1725 2.4062 2.6389 2.871
5 0.00099995 0.0010028 0.0010119 0.0010256 0.0010433 0.0010648 0.0010905 0.42503 0.47443 0.5226 0.57014
10 0.0009997 0.0010026 0.0010117 0.0010254 0.001043 0.0010645 0.0010902 0.206 0.23274 0.25798 0.28249
20 0.00099919 0.0010021 0.0010113 0.0010249 0.0010425 0.0010639 0.0010895 0.0011561 0.11148 0.1255 0.13859
30 0.00099869 0.0010017 0.0010108 0.0010244 0.001042 0.0010633 0.0010888 0.001155 0.070622 0.081175 0.090555
40 0.00099818 0.0010012 0.0010104 0.001024 0.0010415 0.0010628 0.0010881 0.001154 0.0012517 0.058868 0.066474
50 0.00099768 0.0010008 0.0010099 0.0010235 0.001041 0.0010622 0.0010875 0.001153 0.0012499 0.045347 0.051971
60 0.00099718 0.0010003 0.0010095 0.0010231 0.0010405 0.0010616 0.0010868 0.0011521 0.0012481 0.036191 0.042253
70 0.00099668 0.00099987 0.0010091 0.0010226 0.00104 0.0010611 0.0010862 0.0011511 0.0012463 0.029494 0.035265
80 0.00099619 0.00099942 0.0010086 0.0010222 0.0010395 0.0010605 0.0010855 0.0011501 0.0012446 0.02428 0.029978
90 0.00099569 0.00099898 0.0010082 0.0010217 0.001039 0.00106 0.0010849 0.0011491 0.0012429 0.0014024 0.025818
100 0.0009952 0.00099854 0.0010078 0.0010212 0.0010385 0.0010594 0.0010842 0.0011482 0.0012412 0.001398 0.022442
150 0.00099276 0.00099636 0.0010056 0.001019 0.0010361 0.0010567 0.001081 0.0011435 0.001233 0.0013783 0.011481
200 0.00099035 0.0009942 0.0010035 0.0010168 0.0010337 0.001054 0.0010779 0.001139 0.0012254 0.0013611 0.0016649
250 0.00098799 0.00099208 0.0010014 0.0010146 0.0010313 0.0010514 0.0010749 0.0011346 0.0012181 0.0013459 0.0015988
300 0.00098567 0.00098998 0.00099934 0.0010125 0.001029 0.0010488 0.001072 0.0011304 0.0012113 0.0013322 0.0015529
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.5 (continued)
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
350 0.00098338 0.00098792 0.00099731 0.0010104 0.0010267 0.0010462 0.0010691 0.0011264 0.0012048 0.0013197 0.0015175
400 0.00098113 0.00098588 0.00099531 0.0010083 0.0010245 0.0010438 0.0010663 0.0011224 0.0011986 0.0013083 0.0014884
450 0.00097891 0.00098388 0.00099334 0.0010063 0.0010223 0.0010413 0.0010635 0.0011186 0.0011927 0.0012977 0.0014638
500 0.00097673 0.0009819 0.00099139 0.0010043 0.0010201 0.0010389 0.0010608 0.0011149 0.0011871 0.0012879 0.0014424
600 0.00097247 0.00097802 0.00098758 0.0010003 0.0010159 0.0010343 0.0010555 0.0011077 0.0011764 0.00127 0.0014067
Appendix A: Steam Table (Water) According to IAPWS

700 0.00096834 0.00097425 0.00098386 0.00099648 0.0010118 0.0010298 0.0010505 0.001101 0.0011666 0.0012541 0.0013773
800 0.00096434 0.00097057 0.00098025 0.00099276 0.0010078 0.0010255 0.0010456 0.0010945 0.0011574 0.0012398 0.0013525
900 0.00096046 0.00096699 0.00097673 0.00098913 0.001004 0.0010212 0.001041 0.0010884 0.0011488 0.0012268 0.0013309
1000 0.00095669 0.00096351 0.00097329 0.0009856 0.0010002 0.0010172 0.0010364 0.0010826 0.0011407 0.0012148 0.0013118
855
856

m3
Table A.6 Specific volume v of water in kg 2/2
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
0.01 310.6722 333.7492 356.8261 379.9029 402.9796 426.0562 449.1327 472.2092 495.2857
0.1 31.0636 33.372 35.6802 37.9883 40.2963 42.6042 44.9121 47.2199 49.5278
1 3.1027 3.3342 3.5656 3.7968 4.0279 4.259 4.49 4.721 4.952
5 0.61729 0.66421 0.71095 0.75757 0.8041 0.85056 0.89696 0.94333 0.98967
10 0.30659 0.33044 0.35411 0.37766 0.40111 0.4245 0.44783 0.47112 0.49438
20 0.15121 0.16354 0.17568 0.18769 0.19961 0.21146 0.22326 0.23501 0.24674
30 0.099377 0.10788 0.11619 0.12437 0.13244 0.14045 0.1484 0.15631 0.16419
40 0.073432 0.080042 0.086441 0.092699 0.098857 0.10494 0.11097 0.11696 0.12292
50 0.05784 0.063325 0.068583 0.073694 0.078703 0.083637 0.088515 0.09335 0.098151
60 0.047423 0.052168 0.056672 0.061021 0.065264 0.069432 0.073542 0.077609 0.081642
70 0.039962 0.04419 0.048159 0.051966 0.055664 0.059284 0.062847 0.066365 0.069849
80 0.034348 0.038197 0.041769 0.045172 0.048463 0.051673 0.054825 0.057933 0.061005
90 0.029963 0.033528 0.036795 0.039886 0.04286 0.045753 0.048586 0.051374 0.054127
100 0.026439 0.029785 0.032813 0.035655 0.038377 0.041016 0.043594 0.046127 0.048624
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.6 (continued)
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
150 0.015671 0.018478 0.020828 0.022945 0.024921 0.026803 0.028619 0.030387 0.032118
200 0.0099496 0.01272 0.014793 0.016571 0.018184 0.019694 0.021133 0.02252 0.023869
250 0.0060061 0.0091752 0.011142 0.012735 0.01414 0.01543 0.016643 0.017803 0.018922
300 0.0027982 0.0067381 0.0086903 0.010175 0.011444 0.01259 0.013654 0.014662 0.015629
350 0.0021053 0.0049589 0.0069334 0.0083477 0.0095231 0.010566 0.011524 0.012423 0.01328
400 0.0019109 0.0036921 0.0056249 0.0069853 0.0080891 0.0090538 0.009931 0.010748 0.011523
450 0.0018038 0.002915 0.0046344 0.0059384 0.0069828 0.0078848 0.0086979 0.0094507 0.01016
Appendix A: Steam Table (Water) According to IAPWS

500 0.001731 0.0024873 0.0038894 0.0051185 0.0061087 0.0069575 0.0077176 0.0084175 0.0090741
600 0.001633 0.0020847 0.0029518 0.0039548 0.0048336 0.0055908 0.0062651 0.0068818 0.0074568
700 0.0015663 0.0018921 0.0024632 0.0032232 0.0039749 0.0046483 0.0052519 0.0058036 0.0063167
800 0.0015162 0.0017739 0.0021881 0.0027601 0.0033837 0.003975 0.0045161 0.0050133 0.0054762
900 0.0014763 0.0016911 0.0020143 0.0024578 0.0029695 0.0034823 0.0039663 0.0044159 0.004836
1000 0.0014431 0.0016282 0.0018934 0.0022498 0.0026723 0.0031145 0.0035462 0.0039532 0.0043355
857
858

kJ
Table A.7 Specific enthalpy h of water in kg 1/2
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
0.01 −0.041192 2547.5538 2594.4004 2641.3698 2688.536 2735.9492 2783.6463 2879.9959 2977.7344 3076.9529 3177.7157
0.1 −0.032023 104.8447 2591.9936 2639.8034 2687.4309 2735.1334 2783.0201 2879.5902 2977.4457 3076.7343 3177.5435
1 0.059662 104.9281 209.4118 314.0231 2675.7674 2726.6805 2776.5918 2875.4751 2974.5371 3074.5404 3175.8174
5 0.467 105.2985 209.7568 314.3458 419.3985 525.2465 632.2663 2855.8962 2961.1298 3064.5962 3168.0612
10 0.97582 105.7613 210.1879 314.7492 419.7742 525.5915 632.5749 2828.2675 2943.2222 3051.7032 3158.1633
20 1.9923 106.6864 211.0499 315.556 420.5256 526.2821 633.1931 852.5725 2903.2314 3024.2519 3137.6412
30 3.0072 107.6107 211.9116 316.3627 421.2774 526.9734 633.8126 852.9781 2856.5485 2994.3493 3116.0622
40 4.0206 108.5342 212.773 317.1695 422.0295 527.6654 634.4334 853.3874 1085.6861 2961.6515 3093.3182
50 5.0325 109.457 213.634 317.9762 422.7819 528.3581 635.0554 853.8004 1085.662 2925.644 3069.2942
60 6.0429 110.3791 214.4947 318.7829 423.5345 529.0515 635.6786 854.217 1085.6501 2885.4905 3043.8584
70 7.0517 111.3004 215.355 319.5896 424.2875 529.7455 636.3031 854.6371 1085.65 2839.8277 3016.8497
80 8.0591 112.2209 216.215 320.3962 425.0407 530.4402 636.9287 855.0607 1085.6614 2786.3785 2988.061
90 9.0649 113.1408 217.0747 321.2028 425.7942 531.1355 637.5556 855.4876 1085.6839 1344.2693 2957.2186
100 10.0693 114.0599 217.934 322.0094 426.548 531.8315 638.1836 855.9179 1085.7172 1343.0966 2923.9578
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.7 (continued)
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
150 15.0694 118.6443 222.2256 326.0417 430.3208 535.3207 641.3403 858.1171 1086.0356 1338.0633 2692.9998
200 20.0338 123.2107 226.5087 330.0728 434.0996 538.8246 644.5238 860.3911 1086.5836 1334.1395 1645.9511
250 24.9636 127.7595 230.7833 334.1026 437.8839 542.3422 647.7322 862.734 1087.3336 1331.0633 1623.8646
300 29.8599 132.2909 235.0496 338.1307 441.6732 545.8726 650.9638 865.1402 1088.2629 1328.6604 1608.7975
350 34.7235 136.8054 239.3074 342.1569 445.4671 549.415 654.2174 867.605 1089.3525 1326.8077 1597.5402
400 39.5556 141.3033 243.5569 346.1812 449.265 552.9686 657.4915 870.1243 1090.5863 1325.414 1588.7405
450 44.357 145.7849 247.798 350.2032 453.0667 556.5327 660.7849 872.6941 1091.9506 1324.4102 1581.7011
Appendix A: Steam Table (Water) According to IAPWS

500 49.1286 150.2505 252.0309 354.2229 456.8716 560.1065 664.0964 875.3112 1093.4335 1323.7424 1575.9832
600 58.5861 159.1351 260.472 362.2545 464.4903 567.2813 670.7697 880.6748 1096.7157 1323.2514 1567.412
700 67.9346 167.9596 268.8806 370.2752 472.1184 574.4884 677.5039 886.1936 1100.3648 1323.6819 1561.5683
800 77.1804 176.7265 277.2572 378.284 479.7538 581.7238 684.2923 891.8495 1104.3275 1324.8526 1557.6673
900 86.3292 185.438 285.6023 386.2804 487.3947 588.9842 691.1293 897.627 1108.5612 1326.6315 1555.2253
1000 95.386 194.0963 293.9166 394.2638 495.0395 596.2666 698.0098 903.5132 1113.031 1328.9193 1553.9225
859
860

kJ
Table A.8 Specific enthalpy h of water in kg 2/2
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
0.01 3280.0756 3384.0797 3489.7698 3597.1812 3706.3406 3817.2647 3929.9603 4044.4266 4160.6592
0.1 3279.9362 3383.9646 3489.6734 3597.0995 3706.2707 3817.2044 3929.9078 4044.3805 4160.6185
1 3278.5396 3382.8123 3488.7086 3596.2822 3705.5715 3816.6009 3929.3827 4043.9201 4160.2118
5 3272.292 3377.6695 3484.4082 3592.6421 3702.4586 3813.9149 3927.046 4041.8714 4158.4023
10 3264.3855 3371.19 3479.0037 3588.0739 3698.5556 3810.549 3924.1188 4039.3058 4156.1368
20 3248.2271 3358.0519 3468.0932 3578.8756 3690.7089 3803.7889 3918.2436 4034.1585 4151.5935
30 3231.571 3344.6585 3457.0405 3569.5916 3682.8068 3796.9905 3912.3407 4028.9905 4147.0344
40 3214.3735 3330.9912 3445.8374 3560.2183 3674.8479 3790.1538 3906.4104 4023.8023 4142.4601
50 3196.5917 3317.032 3434.4761 3550.7526 3666.8311 3783.2785 3900.4531 4018.5944 4137.8714
60 3178.183 3302.7635 3422.9493 3541.1913 3658.7552 3776.3644 3894.469 4013.3674 4133.2689
70 3159.1044 3288.1695 3411.2503 3531.5318 3650.6193 3769.4116 3888.4587 4008.1219 4128.6531
80 3139.3106 3273.234 3399.3726 3521.7715 3642.4227 3762.4199 3882.4224 4002.8585 4124.0248
90 3118.7526 3257.9419 3387.3105 3511.9082 3634.1646 3755.3896 3876.3607 3997.5778 4119.3844
100 3097.3753 3242.2779 3375.0584 3501.9399 3625.8446 3748.3207 3870.2739 3992.2803 4114.7328
150 2975.5477 3157.8415 3310.7911 3450.474 3583.3076 3712.4081 3839.4829 3965.5633 4091.3257
200 2816.8362 3061.5343 3241.1865 3396.2412 3539.2259 3675.5941 3808.1522 3938.5205 4067.7254
250 2578.7494 2950.3799 3165.9152 3339.2842 3493.6905 3637.973 3776.3704 3911.233 4044.0049
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.8 (continued)
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
300 2152.7882 2820.9128 3084.7924 3279.7892 3446.8724 3599.6773 3744.2417 3883.7843 4020.2341
350 1988.2517 2670.967 2998.0171 3218.0815 3399.0162 3560.8734 3711.8836 3856.2609 3996.4806
400 1931.1915 2511.6302 2906.6872 3154.6483 3350.4327 3521.7571 3679.4249 3828.7509 3972.8094
450 1897.6375 2377.2586 2813.3506 3090.1926 3301.4916 3482.5474 3647.0029 3801.3439 3949.2828
500 1874.3676 2284.3776 2722.5198 3025.7028 3252.6142 3443.4808 3614.7605 3774.1295 3925.9604
600 1843.1916 2179.8228 2570.4044 2902.0646 3156.9527 3366.7648 3551.3945 3720.6353 3880.1539
Appendix A: Steam Table (Water) According to IAPWS

700 1822.9002 2123.3881 2466.3096 2795.013 3067.5056 3293.5699 3490.4519 3668.9601 3835.8142
800 1808.7051 2087.5894 2397.6482 2709.9985 2988.0897 3225.6656 3432.9206 3619.7377 3793.3225
900 1798.473 2062.7176 2350.381 2645.2374 2920.7605 3164.4114 3379.5448 3573.5076 3753.0163
1000 1791.0568 2044.579 2316.2634 2596.0109 2865.0706 3110.6026 3330.7556 3530.6802 3715.1889
861
862

kJ
Table A.9 Specific entropy s of water in kg K 1/2
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
0.01 −0.00015452 9.0921 9.243 9.383 9.5138 9.6368 9.753 9.9682 10.1645 10.3456 10.5142
0.1 −0.00015391 0.36725 8.1741 8.3167 8.4488 8.5725 8.6892 8.9048 9.1014 9.2827 9.4513
1 −0.0001478 0.36723 0.70375 1.0156 7.361 7.4931 7.6147 7.8356 8.0346 8.2171 8.3865
5 −0.00012103 0.36713 0.70357 1.0153 1.3067 1.5813 1.8419 7.0611 7.2726 7.4614 7.6345
10 −8.8423e−05 0.367 0.70334 1.015 1.3063 1.5808 1.8414 6.6955 6.9266 7.1247 7.3028
20 −2.6077e−05 0.36674 0.70287 1.0144 1.3055 1.5798 1.8403 2.3301 6.5474 6.7685 6.9582
30 3.2474e−05 0.36648 0.70241 1.0137 1.3048 1.5789 1.8391 2.3285 6.2893 6.5412 6.7449
40 8.7256e−05 0.36622 0.70195 1.0131 1.304 1.578 1.838 2.3269 2.7933 6.3638 6.5843
50 0.0001383 0.36596 0.70149 1.0125 1.3032 1.577 1.8369 2.3254 2.7909 6.2109 6.4515
60 0.00018563 0.36569 0.70103 1.0119 1.3024 1.5761 1.8358 2.3238 2.7885 6.0702 6.3356
70 0.00022927 0.36543 0.70057 1.0112 1.3017 1.5752 1.8347 2.3223 2.7861 5.9335 6.2303
80 0.00026926 0.36516 0.70011 1.0106 1.3009 1.5743 1.8337 2.3207 2.7837 5.7935 6.1319
90 0.00030561 0.3649 0.69965 1.01 1.3001 1.5734 1.8326 2.3192 2.7814 3.2529 6.0378
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.9 (continued)
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
100 0.00033836 0.36463 0.69919 1.0094 1.2994 1.5724 1.8315 2.3177 2.7791 3.2484 5.9458
150 0.00044893 0.36328 0.69689 1.0063 1.2956 1.5679 1.8262 2.3102 2.7679 3.2275 5.4435
200 0.00047328 0.3619 0.6946 1.0033 1.2918 1.5635 1.8209 2.303 2.7572 3.2087 3.7288
250 0.00041456 0.36051 0.69232 1.0003 1.2881 1.5591 1.8158 2.2959 2.7469 3.1915 3.6803
300 0.00027584 0.35908 0.69004 0.99729 1.2845 1.5548 1.8107 2.289 2.7371 3.1756 3.6435
350 6.0091e−05 0.35764 0.68777 0.99433 1.2809 1.5505 1.8058 2.2823 2.7276 3.1608 3.6131
400 −0.00022982 0.35618 0.68551 0.99139 1.2773 1.5463 1.8009 2.2758 2.7185 3.1469 3.587
450 0.35469 0.68325 0.98848 1.2738 1.5422 1.7961 2.2693 2.7097 3.1338 3.5638
Appendix A: Steam Table (Water) According to IAPWS

−0.00059112
500 −0.0010211 0.35319 0.68099 0.98558 1.2703 1.5381 1.7914 2.2631 2.7012 3.1214 3.543
600 −0.0020771 0.35012 0.67649 0.97987 1.2634 1.5301 1.7822 2.2509 2.6848 3.0982 3.5064
700 −0.0033782 0.34698 0.67201 0.97423 1.2567 1.5223 1.7732 2.2392 2.6694 3.0769 3.4747
800 −0.0049067 0.34377 0.66754 0.96866 1.2501 1.5146 1.7645 2.228 2.6548 3.0572 3.4465
900 −0.0066463 0.34049 0.66309 0.96317 1.2436 1.5071 1.756 2.2171 2.6408 3.0388 3.4211
1000 −0.0085823 0.33716 0.65864 0.95774 1.2373 1.4998 1.7477 2.2066 2.6275 3.0215 3.3978
863
864

kJ
Table A.10 Specific entropy s of water in kgK 2/2
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
0.01 10.6722 10.8212 10.9625 11.0971 11.2258 11.3494 11.4682 11.5829 11.6938
0.1 9.6093 9.7584 9.8997 10.0343 10.1631 10.2866 10.4055 10.5202 10.6311
1 8.5451 8.6945 8.8361 8.9709 9.0998 9.2234 9.3424 9.4571 9.5681
5 7.7954 7.9464 8.0891 8.2247 8.3543 8.4784 8.5977 8.7128 8.824
10 7.4668 7.6198 7.764 7.9007 8.0309 8.1557 8.2755 8.3909 8.5024
20 7.129 7.2863 7.4335 7.5723 7.7042 7.8301 7.9509 8.067 8.1791
30 6.9233 7.0853 7.2356 7.3767 7.5102 7.6373 7.759 7.8759 7.9885
40 6.7712 6.9383 7.0919 7.2353 7.3704 7.4989 7.6215 7.7391 7.8523
50 6.6481 6.8208 6.9778 7.1235 7.2604 7.3901 7.5137 7.6321 7.7459
60 6.5431 6.7216 6.8824 7.0306 7.1692 7.3002 7.4248 7.5439 7.6583
70 6.4501 6.6351 6.7997 6.9505 7.0909 7.2232 7.3488 7.4687 7.5837
80 6.3657 6.5577 6.7264 6.8798 7.0221 7.1557 7.2823 7.403 7.5186
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.10 (continued)
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
90 6.2875 6.4871 6.6601 6.8163 6.9605 7.0955 7.2231 7.3446 7.4608
100 6.2139 6.4217 6.5993 6.7584 6.9045 7.0409 7.1696 7.2918 7.4087
150 5.8817 6.1433 6.3479 6.523 6.6797 6.8235 6.9576 7.0839 7.2039
200 5.5525 5.9041 6.1445 6.339 6.5077 6.6596 6.7994 6.9301 7.0534
250 5.1401 5.6755 5.9642 6.1816 6.3638 6.5246 6.6706 6.8057 6.9324
300 4.4756 5.4419 5.7956 6.0403 6.2374 6.4077 6.5602 6.7 6.8303
350 4.2139 5.1945 5.6331 5.9093 6.1229 6.3032 6.4625 6.6072 6.7411
400 4.1142 4.9446 5.4746 5.7859 6.017 6.2079 6.3743 6.5239 6.6614
Appendix A: Steam Table (Water) According to IAPWS

450 4.0506 4.7361 5.3209 5.6685 5.9179 6.1197 6.2932 6.4479 6.5891
500 4.0029 4.5892 5.1759 5.5566 5.8245 6.0372 6.218 6.3777 6.5226
600 3.9316 4.4134 4.9357 5.3519 5.6528 5.8867 6.0815 6.2512 6.4034
700 3.8777 4.308 4.7663 5.1786 5.5003 5.7522 5.96 6.139 6.2982
800 3.8338 4.2331 4.6475 5.0391 5.3674 5.6321 5.8509 6.0382 6.2039
900 3.7965 4.1748 4.5593 4.9289 5.254 5.5255 5.7526 5.947 6.1184
1000 3.7638 4.1267 4.49 4.8406 5.158 5.4316 5.664 5.8644 6.0405
865
866

kJ
Table A.11 Specific heat capacity c p of water in kg K 1/2
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
0.01 4.2199 1.8732 1.8757 1.8823 1.8913 1.902 1.914 1.9405 1.9693 1.9996 2.0311
0.1 4.2199 4.1822 1.9272 1.9058 1.9057 1.9112 1.9201 1.9436 1.9711 2.0007 2.0318
1 4.2194 4.1819 4.1796 4.1915 2.0741 2.0107 1.9857 1.9757 1.9891 2.0121 2.0396
5 4.2174 4.1807 4.1786 4.1907 4.2157 4.2546 4.3102 2.1448 2.0783 2.0657 2.0753
10 4.215 4.1793 4.1775 4.1896 4.2146 4.2533 4.3086 2.4288 2.2116 2.1408 2.1231
20 4.21 4.1764 4.1752 4.1874 4.2123 4.2506 4.3053 4.4914 2.5602 2.3201 2.2301
30 4.2052 4.1736 4.1729 4.1853 4.21 4.248 4.3022 4.4856 3.0772 2.5431 2.3539
40 4.2003 4.1708 4.1706 4.1831 4.2078 4.2455 4.299 4.4799 4.8646 2.8199 2.4967
50 4.1955 4.168 4.1684 4.181 4.2055 4.2429 4.2959 4.4743 4.8511 3.1714 2.661
60 4.1908 4.1652 4.1661 4.1788 4.2033 4.2403 4.2927 4.4687 4.8379 3.6378 2.8504
70 4.1861 4.1625 4.1639 4.1767 4.2011 4.2378 4.2897 4.4632 4.825 4.2919 3.0704
80 4.1814 4.1597 4.1617 4.1746 4.1989 4.2353 4.2866 4.4578 4.8125 5.287 3.3288
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.11 (continued)
p in ϑ in ◦ C
bar 0 25 50 75 100 125 150 200 250 300 350
90 4.1768 4.157 4.1595 4.1726 4.1967 4.2328 4.2836 4.4525 4.8002 5.7305 3.637
100 4.1723 4.1543 4.1573 4.1705 4.1945 4.2303 4.2806 4.4472 4.7883 5.6816 4.0118
150 4.1501 4.1412 4.1466 4.1603 4.1838 4.2182 4.2659 4.4219 4.7325 5.476 8.7885
200 4.129 4.1285 4.1361 4.1503 4.1734 4.2064 4.2518 4.3982 4.6824 5.3168 8.1062
250 4.109 4.1162 4.126 4.1407 4.1633 4.1951 4.2383 4.3758 4.6371 5.1883 6.98
300 4.0899 4.1044 4.1162 4.1313 4.1534 4.1841 4.2252 4.3547 4.5959 5.0814 6.3935
350 4.0717 4.0931 4.1067 4.1221 4.1439 4.1734 4.2126 4.3347 4.5582 4.9906 6.0154
400 4.0544 4.0821 4.0974 4.1132 4.1346 4.163 4.2005 4.3158 4.5234 4.912 5.7424
Appendix A: Steam Table (Water) According to IAPWS

450 4.038 4.0716 4.0884 4.1046 4.1255 4.153 4.1888 4.2977 4.4912 4.8431 5.5342
500 4.0225 4.0614 4.0797 4.0961 4.1167 4.1432 4.1774 4.2806 4.4613 4.7819 5.37
600 3.9937 4.0422 4.063 4.0798 4.0997 4.1245 4.1558 4.2485 4.4073 4.6775 5.1244
700 3.9678 4.0242 4.0472 4.0644 4.0836 4.1068 4.1355 4.2191 4.3598 4.5912 4.9456
800 3.9446 4.0076 4.0322 4.0497 4.0683 4.09 4.1164 4.1921 4.3174 4.5181 4.8076
900 3.924 3.9921 4.0181 4.0358 4.0537 4.074 4.0984 4.167 4.2794 4.4553 4.6966
1000 3.9057 3.9777 4.0048 4.0225 4.0397 4.0589 4.0813 4.1437 4.2449 4.4003 4.6048
867
868

kJ
Table A.12 Specific heat capacity c p of water in kg K 2/2
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
0.01 2.0635 2.0968 2.1309 2.1656 2.2008 2.2362 2.2716 2.307 2.3423
0.1 2.0641 2.0972 2.1312 2.1659 2.201 2.2364 2.2718 2.3071 2.3424
1 2.0697 2.1015 2.1345 2.1685 2.2031 2.2381 2.2732 2.3083 2.3434
5 2.0952 2.1206 2.1494 2.1803 2.2126 2.2458 2.2795 2.3135 2.3477
10 2.1284 2.1451 2.1682 2.1951 2.2245 2.2555 2.2875 2.3201 2.3532
20 2.1997 2.1964 2.2069 2.2254 2.2486 2.275 2.3035 2.3333 2.3642
30 2.278 2.2508 2.2472 2.2564 2.2732 2.2948 2.3196 2.3467 2.3753
40 2.3642 2.3088 2.2892 2.2883 2.2982 2.3149 2.336 2.3601 2.3865
50 2.459 2.3705 2.333 2.3212 2.3238 2.3353 2.3525 2.3737 2.3978
60 2.5632 2.4364 2.3787 2.355 2.3499 2.3559 2.3692 2.3874 2.4092
70 2.6778 2.5067 2.4265 2.3899 2.3765 2.3769 2.3861 2.4012 2.4206
80 2.8037 2.5817 2.4765 2.4258 2.4038 2.3983 2.4032 2.4152 2.4322
(continued)
Appendix A: Steam Table (Water) According to IAPWS
Table A.12 (continued)
p in ϑ in ◦ C
bar 400 450 500 550 600 650 700 750 800
90 2.9425 2.6617 2.5287 2.4629 2.4316 2.4199 2.4205 2.4293 2.4438
100 3.0958 2.747 2.5833 2.5011 2.46 2.442 2.438 2.4435 2.4555
150 4.1778 3.2687 2.896 2.7112 2.612 2.5575 2.5288 2.5166 2.5155
200 6.3601 4.0074 3.2845 2.9552 2.7812 2.6824 2.6251 2.593 2.5775
250 12.9977 5.086 3.7661 3.2354 2.9679 2.8168 2.7267 2.6725 2.6415
300 25.8263 6.6908 4.3597 3.5529 3.1713 2.9598 2.8331 2.7549 2.7072
350 11.6455 8.9762 5.0715 3.9073 3.3894 3.1103 2.9435 2.8396 2.7744
400 8.7042 10.9505 5.8745 4.2938 3.6194 3.2666 3.057 2.9259 2.8428
Appendix A: Steam Table (Water) According to IAPWS

450 7.4742 10.8642 6.6881 4.7004 3.8571 3.4264 3.1723 3.0133 2.9118
500 6.7794 9.5666 7.309 5.1031 4.0973 3.5873 3.2881 3.1008 2.9813
600 5.9976 7.5398 7.5221 5.7534 4.5557 3.9007 3.5151 3.273 3.1193
700 5.5542 6.5095 6.9698 6.0369 4.9233 4.1825 3.7267 3.4357 3.2526
800 5.2618 5.9182 6.3751 5.9817 5.1372 4.4081 3.9135 3.5834 3.3765
900 5.0515 5.5326 5.9164 5.7779 5.2063 4.5583 4.0692 3.7127 3.4861
1000 4.8915 5.2584 5.576 5.5489 5.1706 4.6275 4.191 3.8235 3.5762
869
Appendix B
Selected Absolute Molar Specific
Enthalpies/Entropies

The following2 tables contain absolute molar specific enthalpies Hm and entropies Sm
as well as specific heat capacities at constant pressure for selected fluids. However,
there are three types of specific heat capacities to be explained in the following:
1. Temperature dependent heat capacities, i.e.

C p = C p (ϑ) (B.1)

2. Arithmetic mean heat capacity


This is useful, for example, to calculate the enthalpy of an ideal gas, i.e.

dHm = C p dT, (B.2)

so that
T

Hm − Hm,0 = C p dT = C p ϑ0 (ϑ − ϑ0 ) . (B.3)
T0

This leads to
T
ϑ 1
C p ϑ0 = C p dT (B.4)
(ϑ − ϑ0 )
T0

Note that Hm,0 = Hm (ϑ0 , T0 ) is a reference level. The enthalpy difference


between two states (1) and (2) then reads as

Hm,2 − Hm,0 = C p ϑ20 (ϑ2 − ϑ0 ) (B.5)

2 The listed data has been taken from NASA Thermo Build, see [45].

© Springer Nature Switzerland AG 2022 871


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
872 Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

and ϑ
Hm,1 − Hm,0 = C p ϑ10 (ϑ1 − ϑ0 ) . (B.6)

Thus, it is ϑ ϑ
Hm,2 − Hm,1 = C p ϑ20 (ϑ2 − ϑ0 ) − C p ϑ10 (ϑ1 − ϑ0 ) . (B.7)

In the following tables, the reference temperature for the averaged heat capacities
is ϑ0 = 0 ◦C.
3. Logarithmic mean heat capacity
This is necessary, for example, to calculate the entropy of an ideal gas. Section 12.3
has shown, that
dT dp
dSm = C p − RM (B.8)
T p

so that

T p
dT dp ϑ T p
Sm − Sm,0 = Cp − RM = C p ϑ0 ln − RM ln . (B.9)
T p T0 p0
T0 p0

This leads to
T
ϑ 1 dT
C p ϑ0 = T Cp (B.10)
ln T0 T
T0

Note that Sm,0 = Sm (ϑ0 , T0 , p0 ) is a reference level. The entropy difference


between two states (1) and (2) then reads as
ϑ T2 p2
Sm,2 − Sm,0 = C p ϑ20 ln − RM ln (B.11)
T0 p0

and
ϑ T1 p1
Sm,1 − Sm,0 = C p ϑ10 ln − RM ln . (B.12)
T0 p0

Thus, it is
   
ϑ2 T2 p2 ϑ1 T1 p1
Sm,2 − Sm,1 
= C p ϑ0 ln − RM ln 
− C p ϑ0 ln − RM ln (B.13)
T0 p0 T0 p0

respectively
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies 873
 
ϑ T2 ϑ T1 p2
Sm,2 − Sm,1 = C p ϑ20 ln − C p ϑ10 ln − RM ln
T0 T0 p1
(B.14)
ϑ2 T2 p
≡ C p ϑ1 ln − RM ln .
2
T1 p1

Thus, it is
ϑ ϑ
ϑ2 C p ϑ20 ln TT20 − C p ϑ10 ln TT01
C p ϑ1 = (B.15)
ln TT21

In the following Tables B.1, B.2, B.3, B.4, B.5, B.6, B.7, B.8, B.9, B.10, B.11,
B.12, B.13, B.14, B.15, B.16, B.17, B.18, B.19, B.20, B.21, B.22, B.23, B.24,
B.25, B.26, B.27, B.28 and B.29 the reference temperature for the averaged heat
capacities is ϑ0 = 0 ◦C.
874

Table B.1 Absolute molar specific enthalpy and entropy of H2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 28.3981 28.5172 28.5157 125.9961 −1.2888
−15 28.4609 28.5463 28.5455 126.5522 −1.1466
−10 28.5196 28.5741 28.5738 127.0987 −1.0042
−5 28.5747 28.6007 28.6007 127.6361 −0.86144
0 28.6262 28.6262 28.6262 128.1644 −0.71844
5 28.6743 28.6505 28.6505 128.6842 −0.57518
10 28.7192 28.6738 28.6735 129.1954 −0.4317
15 28.7611 28.696 28.6954 129.6985 −0.288
20 28.8001 28.7172 28.7162 130.1936 −0.14409
25 28.8363 28.7374 28.7359 130.681 0
50 28.9822 28.8256 28.8207 133.0091 0.72285
100 29.1452 28.9503 28.9371 137.1916 2.1766
150 29.2127 29.0282 29.0078 140.8613 3.6358
200 29.2419 29.0784 29.0527 144.1258 5.0972
250 29.2643 29.1133 29.0836 147.0644 6.5599
300 29.2954 29.1409 29.1076 149.7369 8.0238
350 29.3416 29.166 29.1288 152.1889 9.4897
400 29.4041 29.1917 29.1496 154.4558 10.9583
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.1 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
450 29.4821 29.2195 29.171 156.5652 12.4303
500 29.5742 29.2503 29.1939 158.5393 13.9067
600 29.8007 29.3223 29.2448 162.1494 16.8749
700 30.1045 29.4111 29.3042 165.396 19.8694
800 30.4634 29.5202 29.3741 168.3579 22.8977
900 30.8678 29.6471 29.4526 171.0894 25.9639
1000 31.3025 29.7907 29.5391 173.6319 29.0723
1100 31.7466 29.9483 29.6319 176.0154 32.2247
1200 32.1877 30.1166 29.7293 178.2624 35.4215
1300 32.6185 30.2926 29.8295 180.3904 38.6619
1400 33.0349 30.4737 29.9314 182.4134 41.9447
1500 33.4346 30.6579 30.0339 184.3426 45.2684
1600 33.8167 30.8434 30.1363 186.1875 48.6311
1700 34.1811 31.0291 30.2379 187.9557 52.0311
1800 34.528 31.214 30.3383 189.6542 55.4667
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1900 34.8582 31.3972 30.4372 191.2885 58.9362


2000 35.1726 31.5781 30.5344 192.8638 62.4378
2100 35.4721 31.7565 30.6297 194.3845 65.9702
2200 35.7578 31.9319 30.7231 195.8545 69.5318
2300 36.0308 32.1042 30.8145 197.2773 73.1213
2400 36.2922 32.2733 30.9039 198.656 76.7375
2500 36.543 32.4391 30.9913 199.9935 80.3794
(continued)
875
876

Table B.1 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2600 36.7843 32.6016 31.0767 201.2923 84.0458
2700 37.017 32.7609 31.1601 202.5548 87.736
2800 37.242 32.9169 31.2417 203.7831 91.449
2900 37.4602 33.0699 31.3215 204.9791 95.1841
3000 37.6723 33.2198 31.3995 206.1447 98.9408
3100 37.879 33.3667 31.4758 207.2815 102.7184
3200 38.0811 33.5109 31.5506 208.3911 106.5165
3300 38.2789 33.6524 31.6238 209.4749 110.3345
3400 38.4731 33.7913 31.6955 210.5341 114.1721
3500 38.664 33.9278 31.7658 211.5701 118.029
3600 38.852 34.062 31.8348 212.5839 121.9048
3700 39.0374 34.194 31.9024 213.5767 125.7993
3800 39.2203 34.3239 31.9689 214.5493 129.7122
3900 39.4009 34.4517 32.0342 215.5028 133.6433
4000 39.5792 34.5777 32.0984 216.4379 137.5923
4500 40.4328 35.1815 32.4042 220.8645 157.5982
5000 41.1931 35.7457 32.6873 224.9308 178.0098
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.2 Absolute molar specific enthalpy and entropy of H at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 20.7863 20.7863 20.7863 111.3169 217.0634
−15 20.7863 20.7863 20.7863 111.7235 217.1674
−10 20.7863 20.7863 20.7863 112.1222 217.2713
−5 20.7863 20.7863 20.7863 112.5135 217.3752
0 20.7863 20.7863 20.7863 112.8975 217.4792
5 20.7863 20.7863 20.7863 113.2745 217.5831
10 20.7863 20.7863 20.7863 113.6449 217.687
15 20.7863 20.7863 20.7863 114.0087 217.791
20 20.7863 20.7863 20.7863 114.3663 217.8949
25 20.7863 20.7863 20.7863 114.7179 217.9988
50 20.7863 20.7863 20.7863 116.3916 218.5185
100 20.7863 20.7863 20.7863 119.382 219.5578
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 20.7863 20.7863 20.7863 121.9958 220.5971


200 20.7863 20.7863 20.7863 124.3173 221.6364
250 20.7863 20.7863 20.7863 126.4054 222.6757
300 20.7863 20.7863 20.7863 128.3027 223.7151
350 20.7863 20.7863 20.7863 130.0413 224.7544
(continued)
877
Table B.2 (continued)
878

ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 20.7863 20.7863 20.7863 131.6456 225.7937
450 20.7863 20.7863 20.7863 133.1349 226.833
500 20.7863 20.7863 20.7863 134.5246 227.8723
600 20.7863 20.7863 20.7863 137.0529 229.9509
700 20.7863 20.7863 20.7863 139.3068 232.0296
800 20.7863 20.7863 20.7863 141.34 234.1082
900 20.7863 20.7863 20.7863 143.192 236.1868
1000 20.7863 20.7863 20.7863 144.8923 238.2654
1100 20.7863 20.7863 20.7863 146.464 240.3441
1200 20.7863 20.7863 20.7863 147.9252 242.4227
1300 20.7863 20.7863 20.7863 149.2904 244.5013
1400 20.7863 20.7863 20.7863 150.5714 246.58
1500 20.7863 20.7863 20.7863 151.778 248.6586
1600 20.7863 20.7863 20.7863 152.9185 250.7372
1700 20.7863 20.7863 20.7863 153.9996 252.8158
1800 20.7863 20.7863 20.7863 155.0272 254.8945
1900 20.7863 20.7863 20.7863 156.0064 256.9731
2000 20.7863 20.7863 20.7863 156.9415 259.0517
2100 20.7863 20.7863 20.7863 157.8364 261.1303
2200 20.7863 20.7863 20.7863 158.6944 263.209
2300 20.7863 20.7863 20.7863 159.5183 265.2876
2400 20.7863 20.7863 20.7863 160.3108 267.3662
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.2 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 20.7863 20.7863 20.7863 161.0742 269.4449
2600 20.7863 20.7863 20.7863 161.8106 271.5235
2700 20.7863 20.7863 20.7863 162.5217 273.6021
2800 20.7863 20.7863 20.7863 163.2094 275.6807
2900 20.7863 20.7863 20.7863 163.875 277.7594
3000 20.7863 20.7863 20.7863 164.5199 279.838
3100 20.7863 20.7863 20.7863 165.1455 281.9166
3200 20.7863 20.7863 20.7863 165.7528 283.9953
3300 20.7863 20.7863 20.7863 166.3428 286.0739
3400 20.7863 20.7863 20.7863 166.9165 288.1525
3500 20.7863 20.7863 20.7863 167.4749 290.2311
3600 20.7863 20.7863 20.7863 168.0186 292.3098
3700 20.7863 20.7863 20.7863 168.5484 294.3884
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3800 20.7863 20.7863 20.7863 169.0651 296.467


3900 20.7863 20.7863 20.7863 169.5693 298.5456
4000 20.7863 20.7863 20.7863 170.0615 300.6243
4500 20.7863 20.7863 20.7863 172.3616 311.0174
5000 20.7863 20.7863 20.7863 174.4324 321.4105
879
880

Table B.3 Absolute molar specific enthalpy and entropy of O2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 29.2154 29.2427 29.2423 200.358 −1.3179
−15 29.2279 29.2497 29.2495 200.9295 −1.1718
−10 29.2417 29.2572 29.2571 201.4904 −1.0256
−5 29.257 29.2652 29.2652 202.0409 −0.8794
0 29.2736 29.2736 29.2736 202.5816 −0.73308
5 29.2917 29.2825 29.2825 203.1127 −0.58666
10 29.3112 29.2919 29.2918 203.6348 −0.44016
15 29.3322 29.3018 29.3016 204.148 −0.29355
20 29.3546 29.3122 29.3117 204.6528 −0.14683
25 29.3784 29.323 29.3222 205.1495 0
50 29.5176 29.3841 29.3807 207.5203 0.73613
100 29.8826 29.5379 29.5221 211.7912 2.2207
150 30.3289 29.7254 29.6869 215.5757 3.7257
200 30.8199 29.9371 29.8659 218.9896 5.2543
250 31.3268 30.1643 30.0519 222.1107 6.808
300 31.8283 30.4 30.2394 224.9928 8.3869
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.3 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 32.3098 30.6388 30.4249 227.675 9.9905
400 32.7615 30.8762 30.6055 230.1861 11.6174
450 33.1779 31.1092 30.7795 232.5483 13.2661
500 33.5569 31.3353 30.9459 234.7792 14.9346
600 34.2077 31.7617 31.2537 238.901 18.3239
700 34.7483 32.1507 31.529 242.6397 21.7724
800 35.2233 32.506 31.7764 246.0622 25.2717
900 35.6041 32.8298 31.999 249.2177 28.8137
1000 35.9259 33.1237 32.1992 252.1435 32.3906
1100 36.214 33.3917 32.3805 254.8709 35.9978
1200 36.4828 33.6382 32.546 257.426 39.6328
1300 36.7402 33.867 32.6984 259.8305 43.294
1400 36.9908 34.0812 32.8401 262.1024 46.9806
1500 37.2369 34.2834 32.9727 264.2567 50.692
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1600 37.4798 34.4756 33.0977 266.3063 54.4279


1700 37.7198 34.6594 33.2161 268.2618 58.1879
1800 37.957 34.836 33.3288 270.1325 61.9718
1900 38.1914 35.0065 33.4366 271.926 65.7792
2000 38.4224 35.1715 33.54 273.6494 69.6099
2100 38.65 35.3317 33.6394 275.3084 73.4636
(continued)
881
882

Table B.3 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2200 38.8738 35.4877 33.7354 276.9083 77.3398
2300 39.0934 35.6397 33.8281 278.4535 81.2382
2400 39.3087 35.7881 33.9179 279.9481 85.1583
2500 39.5195 35.9331 34.005 281.3956 89.0998
2600 39.7255 36.0751 34.0896 282.7993 93.0621
2700 39.9266 36.214 34.1718 284.1618 97.0447
2800 40.1228 36.3501 34.2518 285.4859 101.0472
2900 40.314 36.4835 34.3297 286.7737 105.0691
3000 40.5003 36.6143 34.4056 288.0275 109.1099
3100 40.6815 36.7426 34.4797 289.249 113.169
3200 40.8577 36.8685 34.5519 290.4401 117.246
3300 41.029 36.992 34.6225 291.6023 121.3404
3400 41.1955 37.1132 34.6914 292.7371 125.4516
3500 41.3573 37.2321 34.7588 293.8458 129.5793
3600 41.5144 37.3489 34.8247 294.9297 133.7229
3700 41.6669 37.4635 34.8891 295.9899 137.882
3800 41.8149 37.5761 34.9521 297.0275 142.0562
3900 41.9587 37.6867 35.0138 298.0434 146.2449
4000 42.0982 37.7952 35.0742 299.0386 150.4478
4500 42.7369 38.3097 35.3583 303.7326 171.6605
5000 43.2874 38.7806 35.616 308.0177 193.17
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.4 Absolute molar specific enthalpy and entropy of O at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 22.2194 22.1431 22.1441 157.4509 248.1824
−15 22.1801 22.1243 22.1248 157.8851 248.2934
−10 22.1422 22.1059 22.1061 158.3102 248.4042
−5 22.1057 22.088 22.088 158.7266 248.5148
0 22.0704 22.0704 22.0704 159.1347 248.6253
5 22.0364 22.0533 22.0534 159.5347 248.7356
10 22.0035 22.0366 22.0368 159.927 248.8457
15 21.9718 22.0202 22.0207 160.3119 248.9556
20 21.9412 22.0043 22.005 160.6896 249.0654
25 21.9116 21.9887 21.9898 161.0605 249.175
50 21.7781 21.9158 21.9199 162.8193 249.7211
100 21.5698 21.792 21.805 165.937 250.8045
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 21.4184 21.6914 21.715 168.6395 251.879


200 21.3059 21.6084 21.643 171.0251 252.947
250 21.2205 21.539 21.5841 173.161 254.01
300 21.1542 21.4802 21.5351 175.0949 255.0693
350 21.1018 21.4297 21.4938 176.862 256.1257
(continued)
883
Table B.4 (continued)
884

ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 21.0596 21.386 21.4584 178.489 257.1797
450 21.0253 21.3478 21.4277 179.9966 258.2318
500 20.9969 21.3141 21.4009 181.4013 259.2823
600 20.9536 21.2574 21.3563 183.9525 261.3797
700 20.9223 21.2116 21.3206 186.2228 263.4734
800 20.8947 21.1737 21.2911 188.268 265.5642
900 20.8772 21.1416 21.2663 190.1287 267.6528
1000 20.8659 21.1146 21.2453 191.836 269.7399
1100 20.8574 21.0916 21.2273 193.4135 271.826
1200 20.85 21.0718 21.2117 194.8794 273.9114
1300 20.8434 21.0544 21.198 196.2485 275.9961
1400 20.8374 21.0392 21.1859 197.5329 278.0801
1500 20.8324 21.0255 21.175 198.7423 280.1636
1600 20.8285 21.0133 21.1652 199.8852 282.2466
1700 20.8261 21.0024 21.1563 200.9684 284.3293
1800 20.8255 20.9926 21.1482 201.9979 286.4119
1900 20.827 20.9838 21.1409 202.979 288.4945
2000 20.8308 20.976 21.1343 203.9161 290.5774
2100 20.8371 20.9693 21.1283 204.813 292.6607
2200 20.8461 20.9635 21.1229 205.6732 294.7449
2300 20.8577 20.9586 21.1181 206.4998 296.83
2400 20.8721 20.9547 21.1139 207.2953 298.9165
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.4 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 20.8893 20.9517 21.1102 208.0621 301.0046
2600 20.9093 20.9497 21.107 208.8025 303.0945
2700 20.932 20.9486 21.1043 209.5182 305.1865
2800 20.9574 20.9485 21.1021 210.2111 307.281
2900 20.9855 20.9492 21.1004 210.8826 309.3781
3000 21.016 20.951 21.0992 211.5342 311.4781
3100 21.0489 20.9536 21.0984 212.1672 313.5814
3200 21.0842 20.9571 21.098 212.7826 315.688
3300 21.1215 20.9615 21.0981 213.3816 317.7983
3400 21.1609 20.9668 21.0985 213.9652 319.9124
3500 21.2022 20.9729 21.0994 214.5341 322.0305
3600 21.2452 20.9799 21.1006 215.0893 324.1529
3700 21.2897 20.9877 21.1022 215.6314 326.2796
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3800 21.3357 20.9962 21.1041 216.1612 328.4109


3900 21.3829 21.0055 21.1064 216.6792 330.5468
4000 21.4312 21.0156 21.109 217.1861 332.6875
4500 21.6829 21.0756 21.1262 219.5712 343.4655
5000 21.9368 21.1491 21.1491 221.7437 354.3707
885
886

Table B.5 Absolute molar specific enthalpy and entropy of OH at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 30.1067 30.0509 30.0516 178.8324 35.9287
−15 30.0778 30.0372 30.0376 179.421 36.0791
−10 30.0501 30.0239 30.024 179.9977 36.2295
−5 30.0237 30.0109 30.0109 180.5631 36.3796
0 29.9983 29.9983 29.9983 181.1175 36.5297
5 29.974 29.9861 29.9861 181.6615 36.6796
10 29.9507 29.9742 29.9743 182.1953 36.8294
15 29.9284 29.9626 29.9629 182.7194 36.9791
20 29.907 29.9514 29.9519 183.234 37.1287
25 29.8864 29.9404 29.9412 183.7397 37.2782
50 29.7948 29.8898 29.8926 186.1423 38.0242
100 29.6575 29.8057 29.8145 190.4185 39.5103
150 29.566 29.74 29.7556 194.1417 40.9907
200 29.511 29.6889 29.711 197.4405 42.4675
250 29.4885 29.6505 29.678 200.4037 43.9423
300 29.497 29.6238 29.6548 203.0955 45.4168
350 29.5362 29.6081 29.6405 205.564 46.8925
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.5 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 29.6058 29.6032 29.6343 207.8461 48.371
450 29.7051 29.6087 29.6357 209.9707 49.8536
500 29.8329 29.6245 29.644 211.9607 51.342
600 30.1637 29.6855 29.68 215.6082 54.341
700 30.5673 29.7821 29.7378 218.8999 57.3771
800 31.0007 29.907 29.8121 221.9104 60.4553
900 31.4691 30.0543 29.8988 224.6928 63.5786
1000 31.9442 30.2196 29.9947 227.2863 66.7493
1100 32.4074 30.3976 30.0968 229.7191 69.967
1200 32.8496 30.5836 30.2024 232.0127 73.2301
1300 33.2668 30.7742 30.3095 234.1838 76.5361
1400 33.658 30.9663 30.4167 236.246 79.8826
1500 34.0234 31.1581 30.523 238.2105 83.2668
1600 34.3645 31.348 30.6276 240.0865 86.6864
1700 34.6829 31.5349 30.7301 241.8821 90.139
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1800 34.9803 31.7181 30.8301 243.6042 93.6223


1900 35.2587 31.8972 30.9276 245.2586 97.1344
2000 35.5197 32.0719 31.0223 246.8507 100.6735
2100 35.7652 32.242 31.1143 248.3852 104.2378
2200 35.9967 32.4074 31.2036 249.8662 107.826
2300 36.2156 32.5683 31.2903 251.2974 111.4368
2400 36.4233 32.7246 31.3743 252.6821 115.0688
(continued)
887
888

Table B.5 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 36.621 32.8766 31.4559 254.0235 118.7211
2600 36.8096 33.0242 31.5351 255.3241 122.3927
2700 36.9903 33.1678 31.612 256.5866 126.0827
2800 37.1639 33.3074 31.6867 257.8131 129.7905
2900 37.331 33.4433 31.7593 259.0059 133.5153
3000 37.4925 33.5756 31.8299 260.1667 137.2565
3100 37.6488 33.7045 31.8986 261.2973 141.0136
3200 37.8004 33.8301 31.9656 262.3995 144.7861
3300 37.9478 33.9527 32.0308 263.4745 148.5736
3400 38.0913 34.0723 32.0944 264.524 152.3756
3500 38.2313 34.1891 32.1565 265.549 156.1917
3600 38.3678 34.3033 32.2171 266.5508 160.0217
3700 38.5011 34.415 32.2763 267.5306 163.8652
3800 38.6312 34.5242 32.3341 268.4892 167.7218
3900 38.7582 34.6312 32.3907 269.4277 171.5913
4000 38.8821 34.7359 32.4461 270.347 175.4733
4500 39.4505 35.229 32.7061 274.6812 195.0603
5000 39.9103 35.6752 32.941 278.6349 214.9059
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.6 Absolute molar specific enthalpy and entropy of H2 O(liq). The dependency of the enthalpy on the pressure is supposed to be insignificant. Liquid
water is treated as an incompressible fluid, i.e. the entropy does not need to be corrected. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
0 76.1739 76.1739 76.1739 63.3318 −287.717
5 75.6231 75.8626 75.8634 64.7079 −287.3377
10 75.3967 75.6767 75.6789 66.0529 −286.9602
15 75.3332 75.5699 75.5733 67.372 −286.5835
20 75.3354 75.5103 75.5145 68.6679 −286.2068
25 75.351 75.4769 75.4814 69.9422 −285.8301
50 75.3143 75.408 75.4134 76.0085 −283.9466
75 75.5108 75.4004 75.4044 81.6259 −282.062
100 75.9749 75.4745 75.4688 86.875 −280.1696
125 76.7169 75.6272 75.6004 91.8186 −278.2635
150 77.7611 75.8979 75.8294 96.5227 −276.3322
175 78.9244 76.2418 76.1158 101.0172 −274.3745
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

200 80.8047 76.6851 76.4778 105.348 −272.3798


225 83.6512 77.2918 76.9624 109.577 −270.3262
250 87.5755 78.1135 77.6052 113.7633 −268.1885
275 93.3648 79.2164 78.4508 117.975 −265.9324
300 103.521 80.7741 79.6195 122.3399 −263.4846
325 123.3365 83.1999 81.3963 127.1319 −260.6769
889
890

Table B.7 Absolute molar specific enthalpy and entropy of H2 O(g) at p0 = 1 bar. Vapour is treated as an ideal gas. Reference temperature for averaged heat
capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 33.4422 33.4666 33.4663 183.3473 −243.3338
−15 33.4535 33.4729 33.4727 184.0015 −243.1665
−10 33.4658 33.4796 33.4795 184.6434 −242.9993
−5 33.4794 33.4867 33.4867 185.2734 −242.8319
0 33.4942 33.4942 33.4942 185.892 −242.6645
5 33.5103 33.5021 33.5021 186.4998 −242.4969
10 33.5277 33.5105 33.5104 187.0969 −242.3294
15 33.5464 33.5193 33.5191 187.684 −242.1617
20 33.5664 33.5285 33.5281 188.2613 −241.9939
25 33.5877 33.5382 33.5375 188.8291 −241.826
50 33.7132 33.5931 33.59 191.5384 −240.9848
100 34.0481 33.7326 33.7183 196.4108 −239.2912
150 34.4689 33.906 33.8705 200.7174 −237.5786
200 34.9492 34.1057 34.0393 204.5929 −235.8433
250 35.4702 34.326 34.2191 208.1292 −234.083
300 36.0197 34.5621 34.4063 211.3914 −232.2958
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.7 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 36.5904 34.8109 34.5983 214.4275 −230.4806
400 37.1778 35.0699 34.7934 217.2739 −228.6365
450 37.7793 35.3374 34.9907 219.9588 −226.7626
500 38.3934 35.6122 35.1893 222.5048 −224.8584
600 39.6536 36.1802 35.589 227.2495 −220.9563
700 40.943 36.7683 35.9898 231.6176 −216.9266
800 42.229 37.3706 36.3891 235.6845 −212.768
900 43.5057 37.9816 36.7847 239.503 −208.481
1000 44.7428 38.5963 37.1745 243.1121 −204.0682
1100 45.9212 39.2091 37.5564 246.5397 −199.5344
1200 47.0322 39.8152 37.9285 249.8067 −194.8862
1300 48.0727 40.4108 38.2895 252.9298 −190.1304
1400 49.0429 40.9932 38.6387 255.9224 −185.274
1500 49.945 41.5603 38.9756 258.7956 −180.324
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1600 50.782 42.1108 39.3001 261.5588 −175.2872


1700 51.5574 42.644 39.6124 264.2202 −170.1697
1800 52.275 43.1594 39.9125 266.7869 −164.9776
1900 52.9386 43.6568 40.2009 269.2652 −159.7165
2000 53.5521 44.1364 40.4779 271.6608 −154.3916
2100 54.1191 44.5985 40.744 273.9785 −149.0076
2200 54.6431 45.0433 40.9995 276.2232 −143.5692
(continued)
891
892

Table B.7 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 55.1276 45.4714 41.2449 278.3988 −138.0803
2400 55.5758 45.8832 41.4807 280.5092 −132.5449
2500 55.9907 46.2793 41.7074 282.558 −126.9663
2600 56.3752 46.6603 41.9253 284.5483 −121.3477
2700 56.732 47.0268 42.135 286.4832 −115.6922
2800 57.0638 47.3794 42.3368 288.3655 −110.0022
2900 57.373 47.7187 42.5311 290.1978 −104.2802
3000 57.6617 48.0454 42.7183 291.9824 −98.5283
3100 57.9322 48.36 42.8989 293.7218 −92.7484
3200 58.1863 48.6632 43.073 295.418 −86.9424
3300 58.4261 48.9554 43.2412 297.0731 −81.1116
3400 58.6531 49.2373 43.4037 298.6889 −75.2576
3500 58.869 49.5094 43.5608 300.2673 −69.3814
3600 59.0751 49.7723 43.7129 301.8099 −63.4841
3700 59.273 50.0264 43.8601 303.3183 −57.5666
3800 59.4637 50.2723 44.0027 304.794 −51.6298
3900 59.6483 50.5103 44.1411 306.2386 −45.6741
4000 59.8279 50.7411 44.2754 307.6532 −39.7003
4500 60.6762 51.7984 44.8933 314.3201 −9.5715
5000 61.4803 52.7267 45.4378 320.4044 20.9689
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.8 Absolute molar specific enthalpy and entropy of N2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 29.1108 29.1131 29.1131 186.8458 −1.3103
−15 29.1119 29.1137 29.1137 187.4152 −1.1647
−10 29.1131 29.1143 29.1143 187.9737 −1.0191
−5 29.1143 29.115 29.115 188.5217 −0.87357
0 29.1156 29.1156 29.1156 189.0595 −0.72799
5 29.1171 29.1163 29.1163 189.5877 −0.58241
10 29.1186 29.1171 29.1171 190.1065 −0.43682
15 29.1203 29.1179 29.1179 190.6162 −0.29122
20 29.1222 29.1187 29.1187 191.1172 −0.14562
25 29.1244 29.1196 29.1196 191.6097 0
50 29.1389 29.1253 29.125 193.9553 0.72828
100 29.1971 29.1448 29.1428 198.1509 2.1865
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 29.3061 29.1788 29.1724 201.8285 3.6488


200 29.4711 29.2301 29.2151 205.1101 5.118
250 29.6892 29.2993 29.2706 208.081 6.5968
300 29.9526 29.3857 29.3377 210.8025 8.0877
(continued)
893
894

Table B.8 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 30.2512 29.4876 29.4147 213.3198 9.5927
400 30.5741 29.603 29.4998 215.6668 11.1132
450 30.9108 29.7296 29.5911 217.8693 12.6503
500 31.2519 29.8647 29.6867 219.9471 14.2044
600 31.9186 30.152 29.885 223.7885 17.3632
700 32.5384 30.4493 30.0849 227.2829 20.5865
800 33.1019 30.7463 30.2805 230.4933 23.8691
900 33.6008 31.0363 30.4683 233.4649 27.2047
1000 34.0391 31.3152 30.6465 236.2316 30.5872
1100 34.4229 31.5807 30.8145 238.82 34.0107
1200 34.759 31.8318 30.9721 241.2518 37.4702
1300 35.0537 32.0686 31.1197 243.5444 40.9612
1400 35.313 32.2913 31.2579 245.7128 44.4798
1500 35.5418 32.5005 31.3874 247.7694 48.0227
1600 35.7446 32.697 31.5087 249.725 51.5873
1700 35.925 32.8817 31.6225 251.5888 55.1709
1800 36.0863 33.0553 31.7294 253.3689 58.7716
1900 36.2311 33.2187 31.8301 255.0723 62.3876
2000 36.3616 33.3727 31.9249 256.7052 66.0174
2100 36.48 33.5179 32.0144 258.2732 69.6595
2200 36.5876 33.655 32.0991 259.7812 73.313
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.8 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 36.6861 33.7847 32.1793 261.2334 76.9768
2400 36.7765 33.9075 32.2554 262.6339 80.6499
2500 36.8599 34.0239 32.3277 263.9861 84.3318
2600 36.9371 34.1345 32.3965 265.2932 88.0217
2700 37.0089 34.2396 32.4621 266.5582 91.7191
2800 37.0759 34.3398 32.5247 267.7836 95.4233
2900 37.1387 34.4352 32.5846 268.9718 99.1341
3000 37.1977 34.5263 32.6418 270.1251 102.8509
3100 37.2535 34.6134 32.6967 271.2454 106.5735
3200 37.3063 34.6967 32.7494 272.3345 110.3015
3300 37.3565 34.7766 32.7999 273.3942 114.0347
3400 37.4044 34.8532 32.8486 274.426 117.7728
3500 37.4503 34.9267 32.8954 275.4313 121.5155
3600 37.4943 34.9974 32.9406 276.4115 125.2628
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3700 37.5368 35.0655 32.9841 277.3678 129.0143


3800 37.578 35.1311 33.0262 278.3014 132.7701
3900 37.618 35.1943 33.0669 279.2133 136.5299
4000 37.6571 35.2554 33.1062 280.1045 140.2936
4500 37.8455 35.5328 33.2859 284.2818 159.1695
5000 38.0417 35.7737 33.4425 288.0614 178.1404
895
896

Table B.9 Absolute molar specific enthalpy and entropy of N at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 20.7863 20.7863 20.7863 149.9011 471.7446
−15 20.7863 20.7863 20.7863 150.3077 471.8485
−10 20.7863 20.7863 20.7863 150.7064 471.9525
−5 20.7863 20.7863 20.7863 151.0977 472.0564
0 20.7863 20.7863 20.7863 151.4817 472.1603
5 20.7863 20.7863 20.7863 151.8588 472.2643
10 20.7863 20.7863 20.7863 152.2291 472.3682
15 20.7863 20.7863 20.7863 152.5929 472.4721
20 20.7863 20.7863 20.7863 152.9505 472.5761
25 20.7863 20.7863 20.7863 153.3021 472.68
50 20.7863 20.7863 20.7863 154.9758 473.1997
100 20.7863 20.7863 20.7863 157.9662 474.239
150 20.7863 20.7863 20.7863 160.58 475.2783
200 20.7863 20.7863 20.7863 162.9015 476.3176
250 20.7863 20.7863 20.7863 164.9896 477.3569
300 20.7863 20.7863 20.7863 166.887 478.3962
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.9 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 20.7863 20.7863 20.7863 168.6255 479.4355
400 20.7863 20.7863 20.7863 170.2298 480.4749
450 20.7863 20.7863 20.7863 171.7191 481.5142
500 20.7863 20.7863 20.7863 173.1088 482.5535
600 20.7863 20.7863 20.7863 175.6372 484.6321
700 20.7863 20.7863 20.7863 177.891 486.7107
800 20.7806 20.786 20.7861 179.924 488.7891
900 20.7799 20.7853 20.7857 181.7754 490.8671
1000 20.7828 20.7849 20.7855 183.4753 492.9452
1100 20.7865 20.7849 20.7854 185.0469 495.0237
1200 20.7895 20.7851 20.7855 186.5082 497.1025
1300 20.7913 20.7855 20.7857 187.8737 499.1815
1400 20.792 20.786 20.7859 189.155 501.2607
1500 20.7917 20.7864 20.7861 190.362 503.3399
1600 20.7911 20.7867 20.7863 191.5027 505.419
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1700 20.7907 20.7869 20.7864 192.584 507.4981


1800 20.7912 20.7872 20.7865 193.6119 509.5772
1900 20.7935 20.7874 20.7866 194.5913 511.6564
2000 20.7981 20.7878 20.7868 195.5269 513.736
2100 20.8057 20.7885 20.7871 196.4225 515.8161
2200 20.8172 20.7895 20.7876 197.2814 517.8973
2300 20.833 20.791 20.7882 198.1069 519.9797
2400 20.8539 20.7932 20.7891 198.9015 522.064
897

(continued)
898

Table B.9 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 20.8804 20.7961 20.7904 199.6679 524.1507
2600 20.913 20.8 20.792 200.4081 526.2403
2700 20.9522 20.8049 20.794 201.1243 528.3335
2800 20.9984 20.8109 20.7964 201.8181 530.431
2900 21.052 20.8183 20.7994 202.4914 532.5334
3000 21.1134 20.8271 20.8029 203.1455 534.6416
3100 21.1827 20.8374 20.8071 203.7819 536.7564
3200 21.2604 20.8494 20.8118 204.4019 538.8785
3300 21.3464 20.8631 20.8172 205.0066 541.0087
3400 21.441 20.8787 20.8233 205.5971 543.148
3500 21.5442 20.8962 20.8302 206.1743 545.2972
3600 21.656 20.9158 20.8378 206.7393 547.4572
3700 21.7764 20.9374 20.8461 207.2929 549.6287
3800 21.9053 20.9612 20.8553 207.8357 551.8127
3900 22.0426 20.9871 20.8652 208.3687 554.0101
4000 22.188 21.0153 20.876 208.8924 556.2215
4500 23.0267 21.1906 20.9421 211.3915 567.5179
5000 24.009 21.4224 21.0282 213.7327 579.2723
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.10 Absolute molar specific enthalpy and entropy of NO at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 30.0404 29.9887 29.9893 205.8492 89.9241
−15 30.0129 29.9761 29.9764 206.4365 90.0742
−10 29.9874 29.9641 29.9643 207.012 90.2242
−5 29.9638 29.9528 29.9529 207.5762 90.3741
0 29.9422 29.9422 29.9422 208.1295 90.5239
5 29.9224 29.9321 29.9322 208.6725 90.6735
10 29.9046 29.9228 29.9229 209.2054 90.8231
15 29.8887 29.914 29.9142 209.7288 90.9726
20 29.8746 29.9059 29.9063 210.2428 91.122
25 29.8624 29.8984 29.899 210.748 91.2713
50 29.8275 29.8699 29.8715 213.1508 92.0173
100 29.8758 29.8547 29.8565 217.4436 93.5093
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 30.0516 29.888 29.8841 221.21 95.0071


200 30.3212 29.961 29.9437 224.5804 96.516
250 30.6553 30.0656 30.0268 227.6424 98.0402
300 31.0295 30.1946 30.1266 230.4572 99.5822
(continued)
899
900

Table B.10 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 31.4236 30.3419 30.2378 233.0687 101.1435
400 31.8215 30.502 30.3561 235.5091 102.7247
450 32.2108 30.6704 30.4782 237.803 104.3255
500 32.5828 30.8432 30.6014 239.9689 105.9455
600 33.2577 31.1907 30.8443 243.9732 109.2383
700 33.8438 31.5287 31.0753 247.6113 112.594
800 34.3539 31.851 31.2917 250.947 116.0047
900 34.7822 32.1536 31.4922 254.0271 119.4621
1000 35.1458 32.4351 31.6768 256.8874 122.9589
1100 35.4568 32.696 31.8466 259.5568 126.4895
1200 35.7246 32.9375 32.0029 262.0588 130.0489
1300 35.9571 33.161 32.1469 264.4128 133.6332
1400 36.1602 33.3682 32.28 266.6351 137.2393
1500 36.339 33.5604 32.4032 268.7394 140.8644
1600 36.4974 33.7391 32.5176 270.7375 144.5064
1700 36.6386 33.9056 32.6242 272.6394 148.1633
1800 36.7654 34.061 32.7236 274.4539 151.8337
1900 36.88 34.2064 32.8168 276.1886 155.516
2000 36.9841 34.3427 32.9042 277.8502 159.2093
2100 37.0794 34.4708 32.9864 279.4445 162.9125
2200 37.167 34.5914 33.0639 280.9767 166.6249
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.10 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 37.2481 34.7052 33.1371 282.4516 170.3457
2400 37.3237 34.8127 33.2064 283.8732 174.0744
2500 37.3944 34.9146 33.2722 285.2452 177.8103
2600 37.461 35.0113 33.3348 286.5711 181.5531
2700 37.524 35.1032 33.3944 287.8539 185.3024
2800 37.584 35.1907 33.4512 289.0962 189.0578
2900 37.6414 35.2742 33.5056 290.3006 192.8191
3000 37.6967 35.3541 33.5576 291.4694 196.586
3100 37.7501 35.4305 33.6075 292.6047 200.3584
3200 37.802 35.5038 33.6554 293.7083 204.136
3300 37.8528 35.5742 33.7014 294.782 207.9187
3400 37.9027 35.642 33.7458 295.8275 211.7065
3500 37.952 35.7073 33.7886 296.8463 215.4993
3600 38.001 35.7703 33.8299 297.8397 219.2969
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3700 38.0499 35.8312 33.8698 298.809 223.0995


3800 38.0989 35.8903 33.9085 299.7554 226.9069
3900 38.1484 35.9475 33.946 300.6801 230.7192
4000 38.1985 36.0032 33.9824 301.584 234.5366
4500 38.4681 36.2617 34.1505 305.8252 253.7016
5000 38.7953 36.4981 34.301 309.6728 273.0143
901
902

Table B.11 Absolute molar specific enthalpy and entropy of NO2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 35.8377 36.1226 36.119 234.203 32.5508
−15 35.9783 36.1942 36.1922 234.9053 32.7303
−10 36.1212 36.2666 36.2657 235.5969 32.9106
−5 36.2662 36.3396 36.3394 236.2781 33.0915
0 36.4134 36.4134 36.4134 236.9495 33.2732
5 36.5626 36.4878 36.4876 237.6113 33.4557
10 36.7137 36.5629 36.562 238.2641 33.6389
15 36.8665 36.6386 36.6366 238.908 33.8228
20 37.021 36.7148 36.7113 239.5436 34.0075
25 37.177 36.7917 36.7861 240.171 34.193
50 37.9762 37.183 37.1611 243.1961 35.1324
100 39.6257 37.9905 37.9068 248.7749 37.0723
150 41.2704 38.8108 38.6329 253.8593 39.0948
200 42.8502 39.625 39.3281 258.556 41.1982
250 44.3302 40.4199 39.9865 262.9346 43.3782
300 45.693 41.1868 40.6054 267.0432 45.6293
350 46.9326 41.9206 41.1843 270.9169 47.9454
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.11 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 48.0507 42.6182 41.7243 274.5825 50.3205
450 49.0537 43.2786 42.227 278.0615 52.7486
500 49.9506 43.9018 42.6947 281.3712 55.2241
600 51.469 45.0411 43.535 287.5408 60.2979
700 52.6931 46.0498 44.2652 293.1891 65.5081
800 53.6886 46.9446 44.9039 298.3929 70.8289
900 54.508 47.7409 45.4664 303.2133 76.24
1000 55.1909 48.4528 45.9653 307.7006 81.726
1100 55.7674 49.0922 46.411 311.8958 87.2747
1200 56.2615 49.6696 46.8117 315.8336 92.8767
1300 56.6925 50.1936 47.1744 319.543 98.5249
1400 57.0757 50.6717 47.5046 323.0487 104.2136
1500 57.4232 51.1104 47.807 326.3721 109.9388
1600 57.7445 51.5151 48.0856 329.5314 115.6974
1700 58.0474 51.8905 48.3437 332.5425 121.4871
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1800 58.3379 52.2407 48.5839 335.4195 127.3065


1900 58.6207 52.5691 48.8086 338.1743 133.1544
2000 58.8994 52.8786 49.0199 340.8178 139.0305
2100 59.1771 53.1719 49.2194 343.3595 144.9343
2200 59.4559 53.4512 49.4085 345.8077 150.8659
2300 59.7374 53.7184 49.5886 348.1699 156.8256
2400 60.0228 53.9751 49.7606 350.4529 162.8135
(continued)
903
904

Table B.11 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 60.313 54.2228 49.9255 352.6626 168.8303
2600 60.6085 54.4627 50.084 354.8044 174.8763
2700 60.9096 54.6959 50.237 356.8831 180.9522
2800 61.2164 54.9233 50.385 358.9031 187.0584
2900 61.5288 55.1457 50.5284 360.8683 193.1956
3000 61.8466 55.3637 50.6678 362.7823 199.3644
3100 62.1696 55.578 50.8036 364.6483 205.5651
3200 62.4973 55.7891 50.936 366.4694 211.7984
3300 62.8293 55.9974 51.0655 368.2481 218.0647
3400 63.1652 56.2033 51.1922 369.9869 224.3644
3500 63.5044 56.407 51.3164 371.6881 230.6979
3600 63.8464 56.6089 51.4383 373.3537 237.0654
3700 64.1907 56.8092 51.5581 374.9856 243.4672
3800 64.5367 57.008 51.6759 376.5855 249.9036
3900 64.8838 57.2055 51.7918 378.1549 256.3746
4000 65.2317 57.4018 51.906 379.6955 262.8804
4500 66.9647 58.3683 52.4545 387.0081 295.9305
5000 68.6532 59.3129 52.9709 393.7624 329.8375
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.12 Absolute molar specific enthalpy and entropy of CO at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 29.1201 29.123 29.123 192.894 −111.846
−15 29.1214 29.1238 29.1238 193.4636 −111.7004
−10 29.1229 29.1246 29.1246 194.0222 −111.5548
−5 29.1246 29.1256 29.1256 194.5704 −111.4092
0 29.1266 29.1266 29.1266 195.1085 −111.2635
5 29.1288 29.1277 29.1276 195.6368 −111.1179
10 29.1314 29.1289 29.1288 196.1558 −110.9722
15 29.1343 29.1302 29.1301 196.6658 −110.8266
20 29.1376 29.1316 29.1315 197.167 −110.6809
25 29.1414 29.1332 29.1331 197.6598 −110.5352
50 29.1673 29.1432 29.1427 200.0072 −109.8064
100 29.263 29.1766 29.1731 204.2093 −108.3459
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 29.4243 29.2304 29.22 207.8982 −106.879


200 29.6494 29.3057 29.283 211.1963 −105.4024
250 29.9295 29.4017 29.3603 214.1882 −103.9131
300 30.2521 29.516 29.4496 216.9343 −102.4087
350 30.6036 29.646 29.5483 219.4789 −100.8874
(continued)
905
Table B.12 (continued)
906

ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 30.971 29.7886 29.6541 221.8549 −99.3481
450 31.3425 29.9406 29.7645 224.0871 −97.7903
500 31.7085 30.0992 29.8776 226.1947 −96.2139
600 32.3986 30.4258 30.1052 230.0933 −93.008
700 33.0206 30.7528 30.3274 233.64 −89.7366
800 33.5746 31.0718 30.5399 236.8973 −86.4061
900 34.0532 31.3771 30.7402 239.9101 −83.0241
1000 34.4673 31.6659 30.9274 242.7128 −79.5976
1100 34.8258 31.9373 31.1016 245.3327 −76.1325
1200 35.1369 32.1913 31.2636 247.7919 −72.634
1300 35.408 32.4285 31.414 250.1086 −69.1064
1400 35.6451 32.65 31.5539 252.2981 −65.5535
1500 35.8535 32.8568 31.6841 254.3734 −61.9784
1600 36.0377 33.0499 31.8056 256.3455 −58.3836
1700 36.2012 33.2306 31.9191 258.2242 −54.7715
1800 36.3472 33.3997 32.0253 260.0175 −51.144
1900 36.4783 33.5584 32.125 261.7329 −47.5026
2000 36.5965 33.7074 32.2187 263.3767 −43.8487
2100 36.7038 33.8476 32.307 264.9546 −40.1836
2200 36.8017 33.9796 32.3902 266.4715 −36.5083
2300 36.8913 34.1043 32.469 267.9321 −32.8236
2400 36.974 34.2222 32.5436 269.3402 −29.1303
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.12 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 37.0504 34.3338 32.6144 270.6995 −25.429
2600 37.1215 34.4397 32.6818 272.0133 −21.7204
2700 37.1879 34.5402 32.7459 273.2845 −18.0049
2800 37.2502 34.6359 32.807 274.5158 −14.2829
2900 37.3088 34.7271 32.8654 275.7095 −10.5549
3000 37.3642 34.8141 32.9213 276.868 −6.8213
3100 37.4167 34.8972 32.9748 277.9932 −3.0822
3200 37.4667 34.9767 33.0261 279.0871 0.66199
3300 37.5144 35.0529 33.0754 280.1513 4.4111
3400 37.5601 35.126 33.1228 281.1874 8.1648
3500 37.604 35.1962 33.1684 282.1969 11.923
3600 37.6462 35.2636 33.2123 283.1811 15.6855
3700 37.6871 35.3286 33.2548 284.1412 19.4522
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3800 37.7267 35.3912 33.2957 285.0785 23.2229


3900 37.7652 35.4515 33.3353 285.994 26.9975
4000 37.8027 35.5099 33.3736 286.8888 30.7759
4500 37.9822 35.7747 33.5484 291.0817 49.7225
5000 38.1628 36.0044 33.7005 294.8742 68.7582
907
908

Table B.13 Absolute molar specific enthalpy and entropy of CO2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 34.984 35.4743 35.4682 207.8901 −395.1333
−15 35.2305 35.5967 35.5932 208.5767 −394.9578
−10 35.4755 35.7184 35.7169 209.2549 −394.781
−5 35.7188 35.8396 35.8392 209.9249 −394.603
0 35.9601 35.9601 35.9601 210.5871 −394.4238
5 36.1995 36.08 36.0797 211.2415 −394.2434
10 36.4368 36.1992 36.1978 211.8886 −394.0618
15 36.6719 36.3177 36.3145 212.5284 −393.879
20 36.9048 36.4354 36.4298 213.1613 −393.6951
25 37.1354 36.5524 36.5438 213.7874 −393.51
50 38.2528 37.1257 37.0936 216.8223 −392.5675
100 40.3113 38.2135 38.1003 222.4728 −390.6025
150 42.1524 39.2252 38.9993 227.6573 −388.54
200 43.8051 40.1672 39.8081 232.4573 −386.3904
250 45.2963 41.0464 40.5412 236.9326 −384.1622
300 46.648 41.8692 41.2102 241.129 −381.8631
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.13 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 47.8777 42.6411 41.8241 245.0821 −379.4994
400 48.9995 43.3668 42.3902 248.8208 −377.0771
450 50.0246 44.0505 42.9144 252.3683 −374.6011
500 50.9622 44.6955 43.4015 255.7443 −372.0761
600 52.6051 45.8809 44.2797 262.0439 −366.8953
700 53.9797 46.9426 45.0495 267.8233 −361.564
800 55.1417 47.8968 45.7299 273.1608 −356.1063
900 56.1138 48.7574 46.3354 278.1175 −350.5421
1000 56.9323 49.5351 46.8771 282.7417 −344.8887
1100 57.6289 50.2401 47.3645 287.0732 −339.1597
1200 58.2275 50.8814 47.8053 291.1456 −333.3662
1300 58.7459 51.4669 48.2061 294.987 −327.5169
1400 59.1981 52.0033 48.5723 298.6215 −321.6192
1500 59.5951 52.4965 48.9083 302.0696 −315.6791
1600 59.9457 52.9513 49.2179 305.3489 −309.7017
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1700 60.257 53.3721 49.5042 308.4749 −303.6913


1800 60.5347 53.7624 49.7699 311.4608 −297.6514
1900 60.7837 54.1255 50.0172 314.3184 −291.5853
2000 61.0081 54.4641 50.2482 317.0581 −285.4955
2100 61.2112 54.7807 50.4645 319.689 −279.3844
2200 61.396 55.0772 50.6676 322.2193 −273.2539
2300 61.5651 55.3557 50.8587 324.6563 −267.1057
(continued)
909
910

Table B.13 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2400 61.7205 55.6177 51.039 327.0066 −260.9413
2500 61.8643 55.8647 51.2094 329.276 −254.762
2600 61.9982 56.0981 51.3708 331.4699 −248.5688
2700 62.1237 56.319 51.524 333.5932 −242.3626
2800 62.2423 56.5284 51.6697 335.6503 −236.1443
2900 62.3551 56.7274 51.8085 337.6452 −229.9144
3000 62.4633 56.9168 51.9409 339.5817 −223.6734
3100 62.5681 57.0974 52.0675 341.463 −217.4218
3200 62.6703 57.27 52.1888 343.2924 −211.1599
3300 62.771 57.4352 52.305 345.0728 −204.8878
3400 62.8711 57.5936 52.4167 346.8068 −198.6057
3500 62.9713 57.7458 52.5242 348.4969 −192.3136
3600 63.0725 57.8923 52.6277 350.1454 −186.0114
3700 63.1755 58.0337 52.7277 351.7545 −179.699
3800 63.281 58.1704 52.8243 353.3262 −173.3762
3900 63.3898 58.3028 52.9178 354.8623 −167.0427
4000 63.5025 58.4314 53.0084 356.3647 −160.6981
4500 64.15 59.0294 53.4261 363.4253 −128.7917
5000 64.9992 59.5819 53.8004 369.8556 −96.5145
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.14 Absolute molar specific enthalpy and entropy of CGraphite The dependency of the enthalpy on the pressure is supposed to be insignificant. Graphite
is treated as an incompressible solid, i.e. the entropy does not need to be corrected. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 6.9366 7.2972 7.2927 4.4712 −0.34829
−15 7.1178 7.3872 7.3847 4.6086 −0.31315
−10 7.2979 7.4768 7.4757 4.7469 −0.27711
−5 7.477 7.5661 7.5658 4.886 −0.24018
0 7.655 7.655 7.655 5.0257 −0.20235
5 7.8318 7.7435 7.7432 5.1662 −0.16363
10 8.0075 7.8316 7.8306 5.3073 −0.12403
15 8.1821 7.9194 7.917 5.449 −0.083554
20 8.3556 8.0068 8.0027 5.5912 −0.042209
25 8.528 8.0938 8.0874 5.734 9.9811e−09
50 9.3742 8.5235 8.4995 6.4545 0.22383
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

100 10.9941 9.3578 9.271 7.9179 0.73343


150 12.5159 10.1598 9.9826 9.3952 1.3216
200 13.9199 10.927 10.6398 10.8712 1.9831
250 15.1829 11.6544 11.2448 12.3331 2.7113
300 16.2953 12.3372 11.7988 13.7701 3.4988
(continued)
911
Table B.14 (continued)
912

ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm
 J
  J0   J0   J   kJ 
[◦C] mol K mol K mol K mol K mol
350 17.271 12.9739 12.3047 15.1742 4.3385
400 18.1227 13.5655 12.7664 16.5404 5.2238
450 18.865 14.114 13.1882 17.8657 6.149
500 19.5133 14.6222 13.5741 19.1489 7.1088
600 20.5818 15.5302 14.2529 21.5888 9.1158
700 21.4178 16.3139 14.8294 23.8668 11.2174
800 22.086 16.9952 15.3247 25.9951 13.3938
900 22.6321 17.5921 15.755 27.9876 15.6306
1000 23.0882 18.1196 16.1328 29.8578 17.9172
1100 23.4768 18.5894 16.4677 31.6185 20.246
1200 23.8139 19.011 16.7672 33.2808 22.6109
1300 24.1109 19.3921 17.0372 34.8546 25.0074
1400 24.3761 19.7388 17.2822 36.3488 27.432
1500 24.6157 20.0561 17.5061 37.7708 29.8818
1600 24.8347 20.348 17.7118 39.1273 32.3544
1700 25.0371 20.6179 17.9019 40.4243 34.8482
1800 25.2259 20.8688 18.0782 41.6667 37.3614
1900 25.4034 21.1028 18.2426 42.8593 39.893
2000 25.5714 21.3221 18.3964 44.0059 42.4418
2100 25.7314 21.5283 18.5409 45.1103 45.007
2200 25.8843 21.7228 18.677 46.1755 47.5878
2300 26.0313 21.907 18.8057 47.2044 50.1837
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.14 (continued)
ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm
 J
  J0   J0   J   kJ 
[◦C] mol K mol K mol K mol K mol
2400 26.1732 22.0818 18.9277 48.1996 52.7939
2500 26.3106 22.2482 19.0436 49.1634 55.4181
2600 26.4442 22.407 19.154 50.0978 58.0559
2700 26.5745 22.559 19.2594 51.0047 60.7069
2800 26.7018 22.7047 19.3602 51.886 63.3707
2900 26.8266 22.8447 19.4569 52.743 66.0472
3000 26.9492 22.9794 19.5497 53.5773 68.736
3100 27.0698 23.1094 19.639 54.3901 71.4369
3200 27.1887 23.2351 19.7251 55.1827 74.1499
3300 27.3061 23.3567 19.8081 55.9561 76.8746
3400 27.4222 23.4745 19.8884 56.7114 79.611
3500 27.5371 23.589 19.9661 57.4495 82.359
3600 27.6509 23.7002 20.0413 58.1713 85.1184
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3700 27.7638 23.8085 20.1143 58.8776 87.8892


3800 27.8759 23.9141 20.1852 59.5691 90.6712
3900 27.9873 24.0171 20.2541 60.2466 93.4643
4000 28.098 24.1177 20.3211 60.9106 96.2686
4500 28.6444 24.5904 20.6324 64.0496 110.4546
5000 29.1847 25.0229 20.9109 66.9297 124.912
913
914

Table B.15 Absolute molar specific enthalpy and entropy of S at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 23.7146 23.7203 23.7203 163.9527 276.1031
−15 23.7199 23.7213 23.7213 164.4166 276.2217
−10 23.7222 23.7213 23.7213 164.8717 276.3403
−5 23.7217 23.7205 23.7205 165.3182 276.4589
0 23.7188 23.7188 23.7188 165.7564 276.5775
5 23.7136 23.7164 23.7164 166.1866 276.6961
10 23.7063 23.7133 23.7133 166.609 276.8146
15 23.6971 23.7095 23.7096 167.0239 276.9331
20 23.6861 23.705 23.7052 167.4315 277.0516
25 23.6736 23.7 23.7004 167.832 277.17
50 23.5911 23.6674 23.6692 169.7351 277.7609
100 23.367 23.5754 23.5848 173.1139 278.935
150 23.1147 23.4641 23.4869 176.0367 280.0971
200 22.8671 23.3455 23.3863 178.6046 281.2466
250 22.639 23.2266 23.2884 180.8903 282.3842
300 22.436 23.1114 23.1958 182.9474 283.5109
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.15 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 22.2586 23.0019 23.1097 184.8165 284.6282
400 22.105 22.8992 23.0302 186.5284 285.7372
450 21.9721 22.8034 22.9572 188.1074 286.839
500 21.8566 22.7144 22.8901 189.5725 287.9347
600 21.6668 22.5549 22.7718 192.2192 290.1104
700 21.5213 22.4171 22.6711 194.5604 292.2695
800 21.4222 22.2986 22.5852 196.6605 294.4164
900 21.3365 22.1964 22.5115 198.5652 296.5542
1000 21.2665 22.1067 22.4471 200.3077 298.6842
1100 21.2155 22.0279 22.3906 201.9137 300.8082
1200 21.184 21.9587 22.3409 203.4038 302.928
1300 21.1708 21.8985 22.2972 204.7946 305.0456
1400 21.1743 21.8466 22.2589 206.0993 307.1627
1500 21.1925 21.8023 22.2255 207.3289 309.2809
1600 21.2235 21.7651 22.1964 208.4924 311.4016
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1700 21.2654 21.7344 22.1714 209.5973 313.526


1800 21.3165 21.7097 22.1499 210.6498 315.655
1900 21.3752 21.6905 22.1316 211.6554 317.7895
2000 21.44 21.6764 22.1162 212.6184 319.9303
2100 21.5098 21.6668 22.1034 213.5429 322.0777
2200 21.5833 21.6613 22.093 214.4322 324.2323
2300 21.6596 21.6595 22.0846 215.2893 326.3945
2400 21.7377 21.6612 22.0782 216.1165 328.5643
915

(continued)
916

Table B.15 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 21.8169 21.6658 22.0734 216.9163 330.742
2600 21.8965 21.6732 22.0701 217.6906 332.9277
2700 21.9758 21.6829 22.0682 218.4411 335.1213
2800 22.0543 21.6948 22.0675 219.1694 337.3229
2900 22.1316 21.7085 22.0678 219.8768 339.5322
3000 22.2072 21.7239 22.0691 220.5647 341.7491
3100 22.2809 21.7407 22.0712 221.2341 343.9735
3200 22.3522 21.7587 22.074 221.8861 346.2052
3300 22.4209 21.7777 22.0775 222.5215 348.4439
3400 22.487 21.7976 22.0815 223.1413 350.6893
3500 22.5501 21.8182 22.0859 223.7462 352.9412
3600 22.6102 21.8394 22.0908 224.3368 355.1992
3700 22.6673 21.861 22.096 224.9139 357.4631
3800 22.7212 21.8829 22.1015 225.4781 359.7326
3900 22.7719 21.9051 22.1073 226.0298 362.0073
4000 22.8195 21.9273 22.1132 226.5696 364.2869
4500 23.0131 22.0379 22.1444 229.1058 375.7479
5000 23.1466 22.1425 22.176 231.4054 387.2898
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.16 Absolute molar specific enthalpy and entropy of S(a) The dependency of the enthalpy on the pressure is supposed to be insignificant. Sulphur(a)
is treated as an incompressible solid, i.e. the entropy does not need to be corrected. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ C Sm Hm
0
C ϑ C ϑ
 J   J   J0   J   kJ 
[◦C] mol K mol K mol K mol K mol
−40 20.7119 21.384 21.3669 26.7294 −1.4143
−35 20.8907 21.4672 21.4544 27.1708 −1.3102
−30 21.0641 21.5487 21.5396 27.6067 −1.2054
−25 21.2325 21.6287 21.6225 28.0371 −1.0996
−20 21.3961 21.7073 21.7034 28.4623 −0.99304
−15 21.5551 21.7843 21.7822 28.8824 −0.88566
−10 21.71 21.8601 21.8591 29.2973 −0.77749
−5 21.8607 21.9345 21.9343 29.7074 −0.66856
0 22.0076 22.0076 22.0076 30.1126 −0.55889
5 22.1508 22.0795 22.0793 30.5131 −0.44849
10 22.2906 22.1503 22.1494 30.909 −0.33739
15 22.4269 22.2198 22.218 31.3004 −0.22559
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

20 22.56 22.2883 22.2851 31.6874 −0.11313


25 22.69 22.3557 22.3507 32.07 0
50 23.2971 22.6775 22.6594 33.9216 0.57498
75 23.8403 22.9762 22.9392 35.678 1.1643
95.15 24.2371 23.2018 23.1464 37.0305 1.6488
917
918

Table B.17 Absolute molar specific enthalpy and entropy of S(b) The dependency of the enthalpy on the pressure is supposed to be insignificant. Sulphur(b)
is treated as an incompressible solid, i.e. the entropy does not need to be corrected
ϑ C Sm Hm
 J
  J   kJ 
[◦C] mol K mol K mol
95.16 24.7732 38.12 2.05
100 24.8714 38.4441 2.1701
105 24.9729 38.7758 2.2948
110 25.0744 39.1045 2.4199
115.21 25.1801 39.4439 2.5508
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.18 Absolute molar specific enthalpy and entropy of S(liq) The dependency of the enthalpy on the pressure is supposed to be insignificant. Sulphur(liq)
is treated as an incompressible liquid, i.e. the entropy does not need to be corrected
ϑ C Sm Hm
 J
  J   kJ 
[◦C] mol K mol K mol
115.22 31.7109 43.8761 4.2721
150 34.8325 46.6847 5.4116
200 40.222 51.4219 7.5322
250 36.6963 55.2677 9.4449
300 34.9499 58.5299 11.2311
350 33.9564 61.4095 12.9522
400 33.0559 63.997 14.6282
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
919
Table B.19 Absolute molar specific enthalpy and entropy of S2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
Cp
920

ϑ C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 31.5876 31.7999 31.7972 222.9236 127.1575
−15 31.6946 31.8528 31.8513 223.5425 127.3157
−10 31.8006 31.9053 31.9047 224.1515 127.4744
−5 31.9055 31.9574 31.9573 224.751 127.6337
0 32.0091 32.0091 32.0091 225.3414 127.7935
5 32.1114 32.0604 32.0602 225.923 127.9538
10 32.2122 32.1112 32.1106 226.496 128.1146
15 32.3115 32.1614 32.1601 227.0607 128.2759
20 32.4093 32.2112 32.2089 227.6174 128.4377
25 32.5054 32.2605 32.2568 228.1664 128.6
50 32.9611 32.4986 32.4853 230.8021 129.4184
100 33.7483 32.9334 32.8881 235.6012 131.0868
150 34.3854 33.315 33.2281 239.8856 132.7907
200 34.8992 33.6491 33.5165 243.7551 134.5233
250 35.3156 33.9422 33.763 247.2822 136.279
300 35.6571 34.2005 33.9756 250.5216 138.0536
350 35.9424 34.4295 34.1608 253.5161 139.8438
400 36.1867 34.6342 34.3238 256.2997 141.6472
450 36.4019 34.8189 34.4688 258.9002 143.462
500 36.5973 34.9871 34.5994 261.3404 145.287
600 36.9547 35.2857 34.8271 265.8135 148.9649
700 37.2901 35.5483 35.0228 269.8385 152.6773
800 37.6057 35.7858 35.196 273.5013 156.4221
900 37.9326 36.0061 35.3531 276.866 160.1989
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

(continued)
Table B.19 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
1000 38.2636 36.2153 35.4989 279.9823 164.0088
1100 38.5892 36.4164 35.6359 282.8878 167.8515
1200 38.9032 36.6106 35.7656 285.6114 171.7262
1300 39.2019 36.7986 35.8889 288.1763 175.6316
1400 39.483 36.9804 36.0063 290.6009 179.566
1500 39.7454 37.1561 36.1183 292.9005 183.5276
1600 39.9889 37.3256 36.2252 295.0878 187.5145
1700 40.2136 37.489 36.3271 297.1735 191.5247
1800 40.4203 37.6462 36.4245 299.1667 195.5566
1900 40.6098 37.7972 36.5174 301.0753 199.6082
2000 40.7833 37.9423 36.6061 302.9062 203.678
2100 40.9421 38.0814 36.6909 304.6655 207.7644
2200 41.0876 38.2148 36.7719 306.3583 211.866
2300 41.2211 38.3426 36.8494 307.9896 215.9815
2400 41.3442 38.4652 36.9235 309.5636 220.1099
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

2500 41.4582 38.5826 36.9945 311.0842 224.25


2600 41.5646 38.6953 37.0625 312.5547 228.4012
2700 41.6647 38.8034 37.1277 313.9785 232.5628
2800 41.7598 38.9073 37.1904 315.3584 236.734
2900 41.8513 39.0073 37.2506 316.6971 240.9146
3000 41.9402 39.1036 37.3087 317.997 245.1042
3100 42.0278 39.1965 37.3646 319.2605 249.3026
3200 42.115 39.2863 37.4187 320.4896 253.5097
921

(continued)
922

Table B.19 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
3300 42.2029 39.3734 37.471 321.6863 257.7256
3400 42.2922 39.4579 37.5218 322.8524 261.9504
3500 42.3838 39.5402 37.571 323.9896 266.1841
3600 42.4783 39.6205 37.619 325.0995 270.4272
3700 42.5764 39.699 37.6657 326.1836 274.6799
3800 42.6785 39.7761 37.7113 327.2431 278.9426
3900 42.785 39.8519 37.756 328.2796 283.2158
4000 42.8962 39.9266 37.7998 329.294 287.4998
4500 43.5259 40.2904 38.0085 334.0738 309.1003
5000 44.2574 40.6499 38.206 338.4451 331.043
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.20 Absolute molar specific enthalpy and entropy of SO at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J
  J
0   J
0   J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 29.6637 29.7673 29.766 217.0489 3.4144
−15 29.7139 29.7935 29.7928 217.6296 3.5628
−10 29.7661 29.8204 29.82 218.2001 3.7115
−5 29.8201 29.8478 29.8477 218.7609 3.8605
0 29.8758 29.8758 29.8758 219.3123 4.0097
5 29.9332 29.9044 29.9043 219.8547 4.1592
10 29.992 29.9334 29.9331 220.3886 4.3091
15 30.0522 29.963 29.9622 220.9141 4.4592
20 30.1137 29.9929 29.9915 221.4316 4.6096
25 30.1764 30.0233 30.0211 221.9414 4.7603
50 30.5032 30.1807 30.1719 224.3841 5.5188
100 31.1907 30.513 30.4786 228.8204 7.061
150 31.8714 30.8532 30.7796 232.7847 8.6377
200 32.5084 31.1885 31.0661 236.3797 10.2474
250 33.0845 31.5112 31.3336 239.6744 11.8875
300 33.5947 31.8168 31.5809 242.7177 13.5548
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

350 34.0412 32.1034 31.808 245.5464 15.2459


400 34.4302 32.3705 32.0159 248.1889 16.9579
450 34.7695 32.6186 32.2062 250.6681 18.6881
500 35.0674 32.8489 32.3805 253.0027 20.4342
600 35.5675 33.2619 32.6884 257.299 23.9669
700 35.972 33.6212 32.9516 261.1779 27.5445
800 36.322 33.9373 33.1801 264.7137 31.1595
900 36.6203 34.2192 33.3813 267.9631 34.807
923

1000 36.8912 34.473 33.5606 270.9698 38.4827


(continued)
Table B.20 (continued)
ϑ ϑ
Cp Sm Hm
924

ϑ C p ϑ C p ϑ
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
1100 37.1524 34.7047 33.7226 273.769 42.1849
1200 37.4116 34.9195 33.8711 276.3897 45.9131
1300 37.6713 35.1212 34.0087 278.8552 49.6673
1400 37.9315 35.3126 34.1376 281.1847 53.4474
1500 38.191 35.4959 34.2593 283.3941 57.2535
1600 38.4482 35.6724 34.375 285.4964 61.0855
1700 38.7015 35.8431 34.4855 287.5027 64.9431
1800 38.9494 36.0089 34.5913 289.4221 68.8257
1900 39.1906 36.17 34.693 291.2626 72.7327
2000 39.424 36.3269 34.791 293.031 76.6635
2100 39.649 36.4798 34.8855 294.7331 80.6172
2200 39.865 36.6288 34.9768 296.374 84.593
2300 40.0716 36.774 35.065 297.9583 88.5899
2400 40.2688 36.9155 35.1503 299.4899 92.607
2500 40.4565 37.0535 35.2329 300.9722 96.6434
2600 40.6349 37.1878 35.3129 302.4086 100.698
2700 40.8044 37.3186 35.3904 303.8017 104.7701
2800 40.9652 37.446 35.4655 305.1543 108.8586
2900 41.1178 37.57 35.5383 306.4685 112.9628
3000 41.2628 37.6907 35.6089 307.7466 117.0819
3100 41.4006 37.8082 35.6774 308.9904 121.2151
3200 41.5318 37.9225 35.7439 310.2018 125.3618
3300 41.6571 38.0338 35.8085 311.3825 129.5213
3400 41.777 38.1422 35.8713 312.534 133.693
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

(continued)
Table B.20 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
3500 41.8921 38.2477 35.9323 313.6577 137.8765
3600 42.0031 38.3504 35.9916 314.755 142.0713
3700 42.1105 38.4506 36.0494 315.8271 146.277
3800 42.2149 38.5483 36.1056 316.8751 150.4933
3900 42.3169 38.6436 36.1604 317.9002 154.7199
4000 42.4168 38.7367 36.2139 318.9035 158.9566
4500 42.9028 39.1727 36.4631 323.6236 180.287
5000 43.3937 39.5701 36.6878 327.9216 201.8604
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
925
926

Table B.21 Absolute molar specific enthalpy and entropy of SO2 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 38.2268 38.5826 38.5781 241.8393 −298.5664
−15 38.4041 38.6717 38.6692 242.5887 −298.3748
−10 38.5822 38.761 38.7599 243.3271 −298.1824
−5 38.7609 38.8505 38.8502 244.055 −297.989
0 38.9402 38.9402 38.9402 244.7727 −297.7948
5 39.12 39.03 39.0297 245.4807 −297.5996
10 39.3002 39.12 39.1189 246.1793 −297.4036
15 39.4807 39.2102 39.2077 246.8688 −297.2066
20 39.6615 39.3004 39.2961 247.5495 −297.0088
25 39.8425 39.3907 39.3841 248.2218 −296.81
50 40.7466 39.8428 39.8175 251.4658 −295.8026
100 42.5188 40.7407 40.6475 257.4531 −293.7207
150 44.1915 41.6156 41.4233 262.9039 −291.5524
200 45.7273 42.4547 42.1418 267.925 −289.3038
250 47.1103 43.2501 42.8031 272.5882 −286.9822
300 48.3394 43.998 43.4097 276.9448 −284.5954
350 49.4223 44.6972 43.965 281.0335 −282.1507
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.21 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 50.3715 45.3485 44.4729 284.8849 −279.6554
450 51.202 45.9538 44.9379 288.524 −277.1156
500 51.9293 46.5157 45.364 291.9718 −274.5369
600 53.1318 47.5224 46.1155 298.3629 −269.2813
700 54.0741 48.3937 46.7554 304.1762 −263.9192
800 54.8218 49.1522 47.3058 309.5028 −258.473
900 55.4233 49.8167 47.784 314.4144 −252.9597
1000 55.9157 50.4028 48.2033 318.9687 −247.392
1100 56.326 50.9232 48.5742 323.2124 −241.7793
1200 56.6732 51.3882 48.9049 327.1842 −236.1289
1300 56.9713 51.8065 49.202 330.9163 −230.4463
1400 57.2307 52.1849 49.4707 334.4354 −224.7359
1500 57.459 52.5291 49.7151 337.7643 −219.0012
1600 57.6625 52.8437 49.9387 340.9224 −213.2449
1700 57.8457 53.1326 50.1442 343.9262 −207.4694
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1800 58.0125 53.3991 50.3341 346.7901 −201.6763


1900 58.1658 53.646 50.5103 349.5266 −195.8673
2000 58.3081 53.8756 50.6744 352.1467 −190.0435
2100 58.4412 54.0899 50.8277 354.6598 −184.206
2200 58.5669 54.2906 50.9715 357.0745 −178.3555
2300 58.6864 54.4791 51.1068 359.3984 −172.4928
2400 58.801 54.6568 51.2345 361.6381 −166.6184
(continued)
927
928

Table B.21 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 58.9114 54.8248 51.3552 363.7997 −160.7327
2600 59.0185 54.9841 51.4698 365.8885 −154.8362
2700 59.123 55.1354 51.5787 367.9095 −148.9291
2800 59.2255 55.2797 51.6825 369.867 −143.0117
2900 59.3263 55.4175 51.7817 371.7651 −137.0841
3000 59.426 55.5494 51.8766 373.6075 −131.1465
3100 59.525 55.6761 51.9675 375.3973 −125.1989
3200 59.6234 55.7979 52.0549 377.1378 −119.2415
3300 59.7217 55.9153 52.139 378.8316 −113.2742
3400 59.82 56.0287 52.2201 380.4814 −107.2971
3500 59.9185 56.1384 52.2983 382.0895 −101.3102
3600 60.0176 56.2448 52.374 383.6581 −95.3134
3700 60.1172 56.3481 52.4472 385.1893 −89.3067
3800 60.2176 56.4486 52.5182 386.6849 −83.29
3900 60.3189 56.5466 52.5872 388.1467 −77.2631
4000 60.4213 56.6422 52.6542 389.5763 −71.2261
4500 60.9528 57.0913 52.9646 396.2907 −40.8841
5000 61.5262 57.5057 53.2428 402.3906 −10.2664
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.22 Absolute molar specific enthalpy and entropy of CH4 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−20 34.3372 34.5866 34.5833 180.6541 −76.1732
−15 34.4549 34.6503 34.6484 181.3269 −76.0012
−10 34.581 34.7169 34.716 181.989 −75.8286
−5 34.7155 34.7862 34.7859 182.6412 −75.6554
0 34.8582 34.8582 34.8582 183.2838 −75.4815
5 35.0092 34.933 34.9328 183.9175 −75.3068
10 35.1681 35.0105 35.0096 184.5426 −75.1314
15 35.3349 35.0906 35.0885 185.1596 −74.9551
20 35.5094 35.1734 35.1695 185.769 −74.778
25 35.6913 35.2586 35.2526 186.3711 −74.6
50 36.704 35.7213 35.6955 189.2841 −73.6954
100 39.1444 36.8031 36.6917 194.7301 −71.8012
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 41.9488 38.0436 37.7843 199.8222 −69.7749


200 44.9423 39.3916 38.9275 204.6702 −67.6032
250 48.0068 40.8077 40.0898 209.3361 −65.2795
300 51.0666 42.263 41.2508 213.8559 −62.8026
350 54.0755 43.7364 42.3972 218.2516 −60.1737
(continued)
929
Table B.22 (continued)
930

ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
400 57.0066 45.2129 43.521 222.5374 −57.3963
450 59.8452 46.6819 44.6172 226.7229 −54.4746
500 62.5842 48.136 45.6833 230.8151 −51.4135
600 67.7508 50.9804 47.7209 238.7396 −44.8932
700 72.4843 53.7197 49.6318 246.3419 −37.8777
800 76.7676 56.3373 51.4185 253.6414 −30.4116
900 80.6511 58.8267 53.0871 260.6543 −22.5374
1000 84.1528 61.1873 54.6454 267.3955 −14.2942
1100 87.3058 63.4209 56.1012 273.8783 −5.7185
1200 90.1498 65.5318 57.4625 280.116 3.1567
1300 92.7236 67.5261 58.7372 286.1218 12.3025
1400 95.0634 69.4108 59.9328 291.9086 21.6936
1500 97.2011 71.1932 61.0564 297.4894 31.3084
1600 99.1648 72.8809 62.1145 302.8763 41.128
1700 100.9787 74.4811 63.113 308.0812 51.1364
1800 102.6639 76.0005 64.0572 313.1152 61.3195
1900 104.2385 77.4458 64.9521 317.9887 71.6655
2000 105.718 78.8227 65.802 322.7117 82.164
2100 107.1158 80.137 66.6108 327.2931 92.8063
2200 108.4437 81.3938 67.382 331.7417 103.5849
2300 109.7116 82.5976 68.1189 336.0654 114.4931
2400 110.9282 83.7529 68.8243 340.2715 125.5255
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.22 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2500 112.1013 84.8635 69.5008 344.367 136.6773
2600 113.2374 85.9331 70.1507 348.3583 147.9445
2700 114.3422 86.9649 70.776 352.2514 159.3237
2800 115.4209 87.962 71.3788 356.0518 170.8121
2900 116.4779 88.9271 71.9608 359.7646 182.4072
3000 117.5171 89.8628 72.5234 363.3948 194.107
3100 118.542 90.7715 73.0682 366.9468 205.9101
3200 119.5555 91.6552 73.5965 370.4247 217.8151
3300 120.5605 92.5159 74.1094 373.8326 229.8209
3400 121.5592 93.3554 74.6081 377.174 241.9269
3500 122.5538 94.1755 75.0934 380.4525 254.1326
3600 123.5463 94.9775 75.5665 383.6712 266.4376
3700 124.5383 95.7631 76.028 386.8331 278.8419
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3800 125.5314 96.5334 76.4788 389.9411 291.3453


3900 126.5269 97.2897 76.9196 392.9978 303.9482
4000 127.526 98.0331 77.351 396.0057 316.6508
4500 132.6129 101.591 79.3878 410.3918 381.6781
5000 137.9262 104.9566 81.2661 423.8609 449.3016
931
932

Table B.23 Absolute molar specific enthalpy and entropy of C2 H6 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 47.3491 48.44 48.426 221.0732 −86.0957
−15 47.8864 48.7144 48.7065 222.0045 −85.8576
−10 48.4337 48.9919 48.9884 222.9283 −85.6168
−5 48.9904 49.2726 49.2717 223.8452 −85.3733
0 49.5562 49.5562 49.5562 224.7554 −85.1269
5 50.1304 49.8426 49.8417 225.6595 −84.8777
10 50.7125 50.1317 50.1282 226.5578 −84.6256
15 51.302 50.4233 50.4156 227.4507 −84.3705
20 51.8983 50.7174 50.7036 228.3383 −84.1125
25 52.501 51.0138 50.9923 229.2211 −83.8515
50 55.5921 52.5253 52.4407 233.5705 −82.5006
100 61.9938 55.6522 55.3283 242.0156 −79.5617
150 68.4212 58.84 58.1494 250.2078 −76.3009
200 74.6722 62.0217 60.8652 258.1943 −72.7225
250 80.6408 65.1539 63.4572 265.9929 −68.8384
300 86.2822 68.2098 65.9191 273.6098 −64.6639
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.23 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 91.5882 71.1745 68.2523 281.0476 −60.2158
400 96.571 74.0409 70.4618 288.3083 −55.5105
450 101.2523 76.807 72.5552 295.3949 −50.5637
500 105.656 79.474 74.5406 302.3113 −45.3899
600 113.7078 84.5221 78.2187 315.6523 −34.4136
700 120.8042 89.2098 81.5516 328.3682 −22.68
800 127.0036 93.5544 84.5801 340.4891 −10.2833
900 132.4539 97.5805 87.3413 352.049 2.6956
1000 137.2166 101.3114 89.8665 363.0805 16.1845
1100 141.374 104.7687 92.1819 373.6146 30.1187
1200 145.0061 107.9739 94.3105 383.6816 44.4418
1300 148.1856 110.9475 96.2725 393.3107 59.1049
1400 150.9758 113.7092 98.0857 402.5301 74.0659
1500 153.4314 116.2772 99.7656 411.3662 89.2889
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1600 155.5988 118.6684 101.3261 419.8442 104.7426


1700 157.5177 120.8984 102.7791 427.9873 120.4004
1800 159.2218 122.9811 104.1353 435.8173 136.239
1900 160.7396 124.9292 105.404 443.3541 152.2385
2000 162.0956 126.7542 106.5935 450.6164 168.3815
2100 163.3106 128.4666 107.7109 457.6214 184.6529
2200 164.4024 130.0757 108.7629 464.3847 201.0396
(continued)
933
934

Table B.23 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 165.3864 131.5899 109.755 470.921 217.5298
2400 166.2756 133.0169 110.6924 477.2438 234.1137
2500 167.0814 134.3636 111.5796 483.3654 250.7822
2600 167.8137 135.6363 112.4207 489.2974 267.5275
2700 168.4808 136.8406 113.2193 495.0503 284.3427
2800 169.0901 137.9816 113.9788 500.634 301.2217
2900 169.648 139.0641 114.7021 506.0576 318.159
3000 170.1601 140.0922 115.3918 511.3294 335.1498
3100 170.6311 141.0699 116.0504 516.4574 352.1897
3200 171.0655 142.0005 116.68 521.4488 369.2748
3300 171.4667 142.8874 117.2826 526.3103 386.4017
3400 171.8382 143.7335 117.8601 531.0483 403.5672
3500 172.1827 144.5415 118.4141 535.6687 420.7684
3600 172.5027 145.3138 118.946 540.1768 438.0029
3700 172.8006 146.0527 119.4574 544.578 455.2682
3800 173.0781 146.7603 119.9494 548.8768 472.5623
3900 173.3371 147.4385 120.4232 553.0779 489.8832
4000 173.579 148.089 120.8799 557.1854 507.2292
4500 174.5775 150.9797 122.9383 576.45 594.2816
5000 175.3104 153.3779 124.6888 593.8794 681.7626
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.24 Absolute molar specific enthalpy and entropy of C3 H8 at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 64.9696 66.8275 66.8038 259.0084 −107.795
−15 65.8893 67.2939 67.2806 260.2881 −107.4678
−10 66.8204 67.7638 67.7579 261.561 −107.1361
−5 67.7622 68.2372 68.2357 262.8275 −106.7996
0 68.7137 68.7137 68.7137 264.0881 −106.4584
5 69.6741 69.1932 69.1917 265.3432 −106.1125
10 70.6426 69.6755 69.6697 266.5931 −105.7617
15 71.6185 70.1603 70.1474 267.8382 −105.406
20 72.6008 70.6475 70.6246 269.0787 −105.0455
25 73.5888 71.1369 71.1013 270.3149 −104.68
50 78.5904 73.6098 73.4714 276.4383 −102.7779
100 88.6646 78.6217 78.1021 288.4528 −98.5962
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

150 98.4837 83.6172 82.5272 300.2107 −93.9158


200 107.8063 88.511 86.7106 311.7262 −88.7562
250 116.5293 93.2527 90.6426 322.9919 −83.1452
300 124.633 97.8159 94.3286 333.9975 −77.1136
(continued)
935
936

Table B.24 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 132.145 102.1903 97.7821 344.7355 −70.6918
400 139.1159 106.3757 101.0207 355.2034 −63.9082
450 145.6039 110.3781 104.0629 365.4033 −56.7883
500 151.664 114.207 106.9268 375.3404 −49.3549
600 162.6661 121.3863 112.1859 394.4578 −33.6266
700 172.2816 127.987 116.9069 412.6204 −16.8675
800 180.6221 134.0556 121.1647 429.8817 0.78609
900 187.934 139.6452 125.0246 446.3024 19.2222
1000 194.3052 144.7999 128.5388 461.9387 38.3415
1100 199.8548 149.5585 131.7495 476.8427 58.0559
1200 204.6954 153.9561 134.6926 491.0636 78.2889
1300 208.9273 158.0255 137.3988 504.6479 98.9747
1400 212.6373 161.7967 139.8948 517.6394 120.0569
1500 215.8996 165.2971 142.2035 530.0786 141.4872
1600 218.7772 168.5515 144.345 542.0036 163.224
1700 221.3233 171.5824 146.3366 553.4491 185.2316
1800 223.5832 174.4098 148.1936 564.4474 207.4792
1900 225.5954 177.0518 149.9291 575.028 229.94
2000 227.3923 179.5247 151.555 585.2181 252.5911
2100 229.002 181.8432 153.0814 595.0427 275.4122
2200 230.448 184.0202 154.5173 604.5248 298.386
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.24 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 231.7509 186.0676 155.8709 613.6855 321.4971
2400 232.9281 187.996 157.1492 622.544 344.732
2500 233.9947 189.815 158.3585 631.1184 368.079
2600 234.9638 191.5331 159.5045 639.425 391.5277
2700 235.8465 193.1583 160.5923 647.4791 415.0689
2800 236.6526 194.6975 161.6264 655.2946 438.6945
2900 237.3906 196.1571 162.6109 662.8846 462.3972
3000 238.0679 197.543 163.5494 670.2609 486.1706
3100 238.691 198.8604 164.4453 677.4349 510.0089
3200 239.2653 200.1142 165.3017 684.4167 533.9071
3300 239.7959 201.3088 166.1212 691.2159 557.8605
3400 240.2871 202.4481 166.9064 697.8416 581.865
3500 240.7425 203.5358 167.6594 704.3021 605.9168
3600 241.1656 204.5752 168.3824 710.605 630.0124
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

3700 241.5593 205.5696 169.0773 716.7576 654.1489


3800 241.9262 206.5215 169.7457 722.7667 678.3234
3900 242.2685 207.4338 170.3894 728.6387 702.5333
4000 242.5883 208.3087 171.0097 734.3795 726.7764
4500 243.9078 212.1948 173.8048 761.2987 848.418
5000 244.8764 215.4169 176.1808 785.647 970.626
937
938

Table B.25 Absolute molar specific enthalpy and entropy of C2 H5 OH at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 58.7148 60.1209 60.1029 270.4712 −237.7379
−15 59.4084 60.4744 60.4643 271.6263 −237.4425
−10 60.1137 60.8315 60.827 272.7727 −237.1437
−5 60.8298 61.1921 61.191 273.9109 −236.8414
0 61.5559 61.5559 61.5559 275.0414 −236.5354
5 62.2912 61.9229 61.9217 276.1646 −236.2258
10 63.035 62.2926 62.2882 277.281 −235.9125
15 63.7863 62.6651 62.6552 278.3909 −235.5955
20 64.5446 63.0401 63.0225 279.4947 −235.2746
25 65.309 63.4173 63.3899 280.5928 −234.95
50 69.1992 65.3317 65.2246 286.0053 −233.2688
100 77.0956 69.2399 68.835 296.5151 −229.6114
150 84.8097 73.1533 72.3008 306.6879 −225.5624
200 92.1092 76.9899 75.5801 316.5644 −221.1374
250 98.8948 80.7012 78.6578 326.1569 −216.3601
300 105.1476 84.2619 81.5348 335.469 −211.2569
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.25 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 110.894 87.6618 84.2208 344.5038 −205.8538
400 116.1833 90.9008 86.7297 353.267 −200.1751
450 121.0717 93.985 89.0773 361.7666 −194.2422
500 125.6132 96.9235 91.2792 370.013 −188.0737
600 133.8185 102.404 95.3031 385.7918 −175.093
700 140.9635 107.4149 98.8972 400.6919 −161.345
800 147.1451 112.0025 102.126 414.7837 −146.9335
900 152.5829 116.216 105.0455 428.1375 −131.941
1000 157.3368 120.0958 107.6993 440.8152 −116.4397
1100 161.4912 123.6744 110.1217 452.8705 −100.4936
1200 165.1264 126.9806 112.3412 464.3517 −84.1587
1300 168.3145 130.0401 114.3817 475.3025 −67.4832
1400 171.118 132.8763 116.2638 485.7628 −50.5086
1500 173.5906 135.5099 118.0049 495.7687 −33.2707
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1600 175.7778 137.9596 119.6204 505.3532 −15.8


1700 177.7185 140.2424 121.1233 514.5463 1.8767
1800 179.4457 142.3733 122.525 523.3755 19.7366
1900 180.9873 144.3658 123.8357 531.8657 37.7597
2000 182.3674 146.232 125.0639 540.0395 55.9287
2100 183.6063 147.9828 126.2175 547.9176 74.2284
2200 184.7215 149.6278 127.3032 555.5191 92.6458
(continued)
939
940

Table B.25 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 185.7282 151.1759 128.3269 562.8613 111.1691
2400 186.6393 152.6349 129.2941 569.9601 129.7882
2500 187.4659 154.0118 130.2094 576.83 148.4942
2600 188.2179 155.3132 131.077 583.4844 167.2789
2700 188.9037 156.5448 131.9009 589.9358 186.1355
2800 189.5305 157.7118 132.6843 596.1954 205.0577
2900 190.1048 158.8191 133.4303 602.2737 224.0399
3000 190.6322 159.8708 134.1417 608.1806 243.0771
3100 191.1176 160.8711 134.821 613.9249 262.1649
3200 191.5652 161.8234 135.4704 619.5149 281.2994
3300 191.9789 162.731 136.092 624.9586 300.4769
3400 192.3619 163.5969 136.6876 630.2629 319.6941
3500 192.7173 164.4239 137.259 635.4347 338.9483
3600 193.0475 165.2145 137.8077 640.4801 358.2368
3700 193.3548 165.9709 138.3351 645.405 377.557
3800 193.6413 166.6954 138.8426 650.2149 396.907
3900 193.9088 167.3898 139.3313 654.9149 416.2847
4000 194.1589 168.0559 139.8023 659.5096 435.6882
4500 195.1934 171.0167 141.9255 681.0535 533.0399
5000 195.9568 173.4744 143.7313 700.5382 630.8367
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.26 Absolute molar specific enthalpy and entropy of C2 H5 OH(liq). The dependency of the enthalpy on the pressure is supposed to be insignificant.
Liquid ethanol is treated as an incompressible fluid, i.e. the entropy does not need to be corrected. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm

 J
  J
0   J
0   J   kJ 
[ C] mol K mol K mol K mol K mol
−40 94.2886 98.5383 98.4156 135.0857 −284.1454
−35 95.1866 99.0821 98.9859 137.0958 −283.6718
−30 96.157 99.6515 99.5791 139.0835 −283.1934
−25 97.2024 100.2471 100.1957 141.0512 −282.7101
−20 98.3255 100.8695 100.8359 143.0013 −282.2213
−15 99.5284 101.5193 101.5 144.936 −281.7267
−10 100.8135 102.1969 102.1882 146.8575 −281.2258
−5 102.1826 102.9029 102.9007 148.7677 −280.7184
0 103.6376 103.6376 103.6376 150.6688 −280.2039
5 105.1801 104.4015 104.3992 152.5625 −279.6819
10 106.8114 105.1949 105.1854 154.4508 −279.1519
15 108.5328 106.0181 105.9963 156.3353 −278.6136
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

20 110.3454 106.8715 106.832 158.2179 −278.0665


25 112.25 107.7552 107.6924 160.1 −277.51
50 123.1721 112.6355 112.3629 169.5565 −274.5721
75 136.4333 118.2934 117.6347 179.2084 −271.3319
100 151.9748 124.7247 123.4797 189.1894 −267.7314
116 163.0598 129.2379 127.5041 195.7981 −265.2123
941
942

Table B.27 Absolute molar specific enthalpy and entropy of CH3 OH at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 41.5565 42.0448 42.0385 232.8241 −202.8629
−15 41.7917 42.1687 42.1651 233.6392 −202.6546
−10 42.0378 42.2962 42.2945 234.4432 −202.445
−5 42.2944 42.4271 42.4267 235.2368 −202.2342
0 42.5615 42.5615 42.5615 236.0207 −202.022
5 42.8385 42.6992 42.6988 236.7952 −201.8085
10 43.1253 42.8401 42.8385 237.5609 −201.5936
15 43.4213 42.9843 42.9804 238.3184 −201.3773
20 43.7261 43.1314 43.1246 239.068 −201.1594
25 44.0395 43.2816 43.2708 239.8101 −200.94
50 45.7184 44.073 44.0287 243.4217 −199.8184
100 49.4777 45.8184 45.6382 250.2579 −197.4402
150 53.4971 47.7046 47.3035 256.7257 −194.8664
200 57.5467 49.66 48.9674 262.9229 −192.09
250 61.4901 51.6342 50.5953 268.8999 −189.1135
300 65.2553 53.5934 52.1675 274.6834 −185.944
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.27 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 68.8119 55.516 53.6744 280.2895 −182.5914
400 72.1547 57.3891 55.1124 285.7291 −179.0664
450 75.2924 59.2059 56.4819 291.0111 −175.3794
500 78.2398 60.9634 57.7852 296.1434 −171.5403
600 83.6226 64.3004 60.2078 305.9873 −163.4418
700 88.3737 67.4076 62.4097 315.3132 −154.8367
800 92.5326 70.2934 64.4149 324.1616 −145.7873
900 96.2113 72.9731 66.2469 332.5707 −136.3462
1000 99.4471 75.4622 67.926 340.5743 −126.5598
1100 102.2884 77.7746 69.4691 348.2022 −116.47
1200 104.7846 79.9237 70.8907 355.4812 −106.1136
1300 106.9811 81.9223 72.2038 362.436 −95.523
1400 108.9183 83.7829 73.4196 369.0894 −84.726
1500 110.6312 85.5168 74.5481 375.4623 −73.7468
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1600 112.1499 87.1349 75.5982 381.5741 −62.6063


1700 113.5003 88.6468 76.5775 387.4424 −51.3225
1800 114.7043 90.0616 77.493 393.0837 −39.9111
1900 115.7808 91.3875 78.3506 398.5129 −28.3858
2000 116.746 92.6317 79.1557 403.7437 −16.7586
2100 117.6136 93.801 79.913 408.7887 −5.0399
2200 118.3956 94.9015 80.6266 413.6594 6.7612
(continued)
943
944

Table B.27 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2300 119.1023 95.9386 81.3004 418.3665 18.6367
2400 119.7425 96.9173 81.9377 422.9199 30.5795
2500 120.324 97.8421 82.5414 427.3283 42.5833
2600 120.8533 98.7171 83.1142 431.6003 54.6425
2700 121.3364 99.5461 83.6585 435.7434 66.7524
2800 121.7782 100.3323 84.1765 439.7647 78.9084
2900 122.1832 101.0789 84.6702 443.6708 91.1068
3000 122.5553 101.7887 85.1412 447.4677 103.344
3100 122.8979 102.4642 85.5912 451.1611 115.6169
3200 123.214 103.1077 86.0217 454.7562 127.9227
3300 123.5062 103.7215 86.4339 458.2579 140.2589
3400 123.7768 104.3074 86.8291 461.6707 152.6232
3500 124.0279 104.8673 87.2084 464.9988 165.0136
3600 124.2613 105.4029 87.5727 468.2462 177.4282
3700 124.4787 105.9155 87.9231 471.4166 189.8653
3800 124.6813 106.4067 88.2603 474.5133 202.3235
3900 124.8706 106.8777 88.5852 477.5397 214.8012
4000 125.0476 107.3298 88.8984 480.4988 227.2972
4500 125.7806 109.3413 90.3113 494.3779 290.0138
5000 126.3225 111.0136 91.5143 506.9362 353.0461
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.28 Absolute molar specific enthalpy and entropy of CH3 OH(liq). The dependency of the enthalpy on the pressure is supposed to be insignificant.
Liquid methanol is treated as an incompressible fluid, i.e. the entropy does not need to be corrected. Reference temperature for averaged heat capacities is
ϑ0 = 0 ◦C
ϑ ϑ
ϑ C C ϑ C ϑ Sm Hm
 J
  J
0   J
0   J   kJ 
[◦C] mol K mol K mol K mol K mol
−40 72.5503 74.5339 74.476 108.5685 −243.8644
−35 72.958 74.7886 74.7431 110.1121 −243.5007
−30 73.4066 75.0569 75.0226 111.6326 −243.1348
−25 73.8958 75.3387 75.3143 113.1317 −242.7665
−20 74.4254 75.634 75.618 114.611 −242.3958
−15 74.9956 75.943 75.9338 116.0722 −242.0222
−10 75.6065 76.2657 76.2615 117.5166 −241.6457
−5 76.2587 76.6022 76.6011 118.9458 −241.2661
0 76.9527 76.9527 76.9527 120.3609 −240.8831
5 77.6894 77.3175 77.3164 121.7634 −240.4965
10 78.4694 77.6966 77.692 123.1544 −240.1061
15 79.2938 78.0904 78.0799 124.5351 −239.7117
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

20 80.1637 78.499 78.4801 125.9066 −239.3131


25 81.08 78.9228 78.8927 127.27 −238.91
50 86.3963 81.2782 81.1466 134.0013 −236.8192
75 93.0188 84.0505 83.7289 140.6746 −234.5793
100 100.9939 87.2616 86.6454 147.3908 −232.1569
116 106.7768 89.5487 88.683 151.7497 −230.4954
945
946

Table B.29 Absolute molar specific enthalpy and entropy of air at p0 = 1 bar. Reference temperature for averaged heat capacities is ϑ0 = 0 ◦C
ϑ ϑ
ϑ Cp Sm Hm
0
C p ϑ C p ϑ
 J   J   J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
−20 29.0566 29.0643 29.0642 194.0651 −1.434
−15 29.0602 29.0663 29.0662 194.6335 −1.2887
−10 29.0641 29.0684 29.0684 195.191 −1.1434
−5 29.0683 29.0706 29.0706 195.7381 −0.99805
0 29.0729 29.0729 29.0729 196.2752 −0.85269
5 29.0779 29.0754 29.0754 196.8026 −0.70732
10 29.0833 29.078 29.0779 197.3207 −0.56191
15 29.0891 29.0807 29.0806 197.8298 −0.41648
20 29.0953 29.0836 29.0834 198.3303 −0.27102
25 29.102 29.0866 29.0864 198.8224 −0.12553
50 29.1429 29.104 29.103 201.1672 0.60251
100 29.2654 29.1517 29.1468 205.3678 2.0625
150 29.4447 29.218 29.2048 209.0582 3.53
200 29.6769 29.3026 29.2759 212.3591 5.0078
250 29.9538 29.4045 29.3584 215.3536 6.4984
300 30.265 29.5216 29.4503 218.1015 8.0038
(continued)
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Table B.29 (continued)
ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
350 30.5994 29.6515 29.5495 220.6466 9.5253
400 30.9465 29.7916 29.6539 223.0214 11.0639
450 31.297 29.9394 29.7618 225.2511 12.62
500 31.643 30.0925 29.8715 227.355 14.1936
600 32.3004 30.4065 30.0911 231.2436 17.3912
700 32.8981 30.7204 30.3051 234.7783 20.6516
800 33.438 31.0271 30.5099 238.0228 23.969
900 33.9076 31.3217 30.7034 241.023 27.3368
1000 34.3176 31.6012 30.8846 243.8137 30.7485
1100 34.6778 31.8649 31.0539 246.4222 34.1987
1200 34.9967 32.1129 31.2118 248.8712 37.6827
1300 35.281 32.3458 31.3591 251.1791 41.1969
1400 35.536 32.5647 31.4968 253.3613 44.738
1500 35.7664 32.7706 31.6258 255.4309 48.3033
1600 35.9757 32.9645 31.7468 257.399 51.8905
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies

1700 36.167 33.1474 31.8605 259.2751 55.4978


1800 36.3427 33.3201 31.9677 261.0675 59.1234
1900 36.5049 33.4835 32.069 262.7834 62.7659
2000 36.6553 33.6383 32.1647 264.4291 66.424
2100 36.7955 33.7854 32.2556 266.0102 70.0966
2200 36.9265 33.9252 32.3419 267.5316 73.7828
2300 37.0494 34.0584 32.424 268.9977 77.4816
(continued)
947
948

Table B.29 (continued)


ϑ ϑ
ϑ Cp C p ϑ C p ϑ Sm Hm
 J   J 0  J 0  J   kJ 
[◦C] mol K mol K mol K mol K mol
2400 37.1651 34.1855 32.5023 270.4125 81.1924
2500 37.2744 34.3069 32.577 271.7795 84.9144
2600 37.3779 34.423 32.6485 273.1018 88.6471
2700 37.4761 34.5343 32.717 274.3823 92.3899
2800 37.5696 34.641 32.7827 275.6236 96.1422
2900 37.6587 34.7436 32.8458 276.828 99.9036
3000 37.7439 34.8422 32.9064 277.9978 103.6738
3100 37.8254 34.9371 32.9648 279.1349 107.4523
3200 37.9036 35.0286 33.0211 280.2412 111.2388
3300 37.9787 35.1169 33.0754 281.3181 115.0329
3400 38.051 35.2021 33.1279 282.3674 118.8344
3500 38.1208 35.2845 33.1786 283.3904 122.643
3600 38.1881 35.3642 33.2277 284.3885 126.4585
3700 38.2533 35.4414 33.2753 285.3628 130.2806
3800 38.3165 35.5163 33.3213 286.3144 134.1091
3900 38.3778 35.5888 33.366 287.2445 137.9438
4000 38.4376 35.6593 33.4095 288.154 141.7846
4500 38.7187 35.9839 33.6094 292.4228 161.0747
5000 38.9875 36.2708 33.7858 296.2932 180.5013
Appendix B: Selected Absolute Molar Specific Enthalpies/Entropies
Appendix C
Caloric State Diagrams

C.1 Water

See Figs. C.1, C.2 and C.3.

© Springer Nature Switzerland AG 2022 949


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
950 Appendix C: Caloric State Diagrams

Fig. C.1 log p, h-diagram of water, generated with CoolProp, see [8]
Appendix C: Caloric State Diagrams 951

Fig. C.2 T, s-diagram of water, generated with XSteam, see [9]


952 Appendix C: Caloric State Diagrams

Fig. C.3 h, s-diagram of water, generated with XSteam, see [9]


Appendix C: Caloric State Diagrams 953

C.2 Refrigerants

See Figs. C.4, C.5, C.6, C.7 and C.8.


954 Appendix C: Caloric State Diagrams

Fig. C.4 log p, h-diagram of R134a, generated with CoolProp, see [8]
Appendix C: Caloric State Diagrams 955

Fig. C.5 log p, h-diagram of R290, generated with CoolProp, see [8]
956 Appendix C: Caloric State Diagrams

Fig. C.6 log p, h-diagram of R717, generated with CoolProp, see [8]
Appendix C: Caloric State Diagrams 957

Fig. C.7 log p, h-diagram of R744, generated with CoolProp, see [8]
958 Appendix C: Caloric State Diagrams

Fig. C.8 log p, h-diagram of R1234yf, generated with CoolProp, see [8]
Appendix D
The h1+x , x-Diagram

See Figs. D.1 and D.2.

© Springer Nature Switzerland AG 2022 959


A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
960 Appendix D: The h 1+x , x-Diagram

Fig. D.1 h 1+x , x-diagram (atmospheric air + water)


Appendix D: The h 1+x , x-Diagram 961

Fig. D.2 h 1+x , x-diagram (hydrogen + water)


Technical Thermodynamics

State values with Ėx,S,i = ṁex,S,i


 
Z δq δψ 2 1
z= h = u + pv ds = + = δsa + δsi = ṁ hi − henv + c2i + gzi + Tenv (senv − si )
2
m T T W12 = W12,V + W12,mech + Ψ12 and W12,V = − p dV
 1

1 2 Tenv
Δea,12 = c − c21 + g (z2 − z1 ) ΔĖx,V = Tenv Ṡi and Ėx,Q = Q̇12 1 −
2 2 Weff = W12 + penv (V2 − V1 ) Tm
Equations of state (ideal gas) Open system - unsteady system
2. Law
1. Law
pV = mRT resp. pv = RT
Appendix E
Formulary

 Q12
S2 − S1 = m (s2 − s1 ) = Sa,12 + Si,12 = + Si,12 
Tm c2Sys,Δt
mSys,Δt uSys,Δt + + gzSys,Δt +
2
c =const.
v Exergy 
du = cv dT −− −−−→ u2 − u1 = cv (T2 − T1 ) c2Sys,0
− mSys,0 uSys,0 + + gzSys,0 =
Ex,2 = Ex,1 + Weff + Ex,Q − ΔEx,V with ΔEx,V = Tenv Si,12 2
cp =const.
dh = cp dT −−−−−→ h2 − h1 = cp (T2 − T1 ) Δt
 c2
Tenv h1 + 1 + gz1 ṁ1 dt+
du p dv dh v dp Ex,Q = Q12 1 − = Q12 − Tenv (S2 − S1 ) + Tenv Si,12 2
ds = + = − Tm 0
T T T T
Δt Δt Δt
cv , cp =const. T2 p2 T2 v2 c22
−−−−−−−→ s2 − s1 = cp ln − R ln = cv ln + R ln Ex,i = mi (ui − uenv ) + − h2 + + gz2 ṁ2 dt + Q̇ dt + Pt dt
T1 p1 T1 v1   2

© Springer Nature Switzerland AG 2022


1 0 0 0
+ mi c2i + gzi + Tenv (senv − si ) − penv (venv − vi )
2

https://doi.org/10.1007/978-3-030-97150-2
Thermal flow machines
ϑ ϑ
cp ϑ c 2◦ ϑ2 − c0 1◦ C ϑ1 Open system - steady state flow process
cp = cv + R κ= = 0 C h2 − h1 h2s − h1
cϑ21 Continuity ηs,T = and ηs,V =
cv ϑ2 − ϑ1 h2s − h1 h2 − h1
ϑ ϑ   cA  κ−1
cv  2 ln T2 − cv  1 ln T1 p2 κ
ϑ ϑ0 T0 ϑ0 T0 ṁi = ṁj with ṁ = ρcA =
cv ϑ21 = i,in j,out
v T2s = T1
p1

A. Schmidt, Technical Thermodynamics for Engineers,


ln TT21
kJ m Thermodynamic cycle
1. Law
RM = R · M = 8.3143 M= 
kmol K n     
 1
 
1 dZ = 0 and Q+ W =0
Equations of state (incompressible liquids) Q̇ + Pt + ṁi hi + c2i + gzi = ṁj hj + c2j + gzj
i,in
2 j,out
2
c=const.
du = c dT −−−−−→ u2 − u1 = c (T2 − T1 )
with
dh = c dT + v dp Tmin ex. benefit
2   ηC = 1 − and ηex =
c=const. Pt 1 2  Tmax ex. effort
−−−−−→ h2 − h1 = c (T2 − T1 ) + v (p2 − p1 ) wt,12 = = v dp +ψ12 + c − c21 + g (z2 − z1 )
ṁ 2 2 ηth = ηex ηC
du c=const. T2
1
ds = −−−−−→ s2 − s1 = c ln =y12 benefit P
T T1 ηth,TE = =
effort Q̇in
Thermodynamic mean temperature 2. Law

q12 + ψ12 q12 ψ12    Q̇


Tm = Tm = Tm = Ṡi + Ṡa + ṁi si = ṁj sj with Ṡa =
s2 − s1 sa,12 si,12 Tm benefit Q̇in Tmin
i,in j,out εCM = = εCM,C =
effort P Tmax − Tmin
Closed system
1. Law Exergy
benefit Q̇out Tmax
    εHP = = εHP,C =
1 2  effort P Tmax − Tmin
W12 + Q12 = U2 − U1 + m c − c21 + g (z2 − z1 ) ṁi ex,S,i + Ėx,Q + Pt = ΔĖx,V + ṁj ex,S,j
2 2 i,in j,out εHP = εCM + 1

963
964

Technical Thermodynamics

isochoric isobaric isothermal isentropic polytropic

p,v

T,s

p1 v1κ = p2 v2κ p1 v1n = p2 v2n


p1 p2 v1 v2  κ−1  n−1
p,v,T = = p1 v 1 = p 2 v2
T1 T2 T1 T2 T2 p2 κ T2 p2 n
= =
T1 p1 T1 p1

T2 T2 v2
s2 − s1 = cv ln s2 − s1 = cp ln s2 − s1 = R ln
T1 T1 v1 n − κ T2
s 2 − s1 p2 v2 p1 s 2 − s1 = 0 s2 − s1 = cv ln
s2 − s1 = cv ln s2 − s1 = cp ln s2 − s1 = R ln n − 1 T1
p1 v1 p2

 κ−1 
p1 v1 v1  n−1 
v2 w12,V = −1
w12,V = −RT ln p 1 v1 v1
v1 κ−1 v2 w12,V = −1
p1   n−1 v2
2 w12,V = −RT ln RT1 T2  
w12,V = −p (v2 − v1 ) p2 w12,V = −1 RT1 T2
w12,V w12,V = − p dv = 0 p1 κ − 1 T1 w12,V = −1
w12,V = −R (T2 − T1 ) w12,V = −p1 v1 ln n − 1 T1
1 p2 w12,V = cv (T2 − T1 )  n−1 
 κ−1  RT1 p2 n
p1 RT1 p2 κ w12,V = −1
w12,V = −p2 v2 ln w12,V = −1 n−1 p1
p2 κ−1 p1

2
y12 y12 = v dp = v (p2 − p1 ) y12 = 0 y12 = w12,V y12 = κw12,V y12 = nw12,V
1
Appendix E: Formulary

2
Technical Thermodynamics

Real fluids Humid air ha = cp,a · ϑ hv = Δhv,0 + cp,v · ϑ hice = Δhm,0 + cice · ϑ
 
mass vavpour mV mw mv Ra p v pv Δhm,0 T
x= = x= xv = = = 0.622 s1+x = s1+x + (x − x ) · + cice ln
mass total mV + m L ma ma Rv p − p v p − pv T0,w T0,w
v − v u − u h − h s − s Z
x= =  =  =  z1+x =
v  − v  u − u h − h s − s ma
J J
Δhv = h − h Ra = 287.1 Rv = 461.5
kg K kg K
Clausius-Clapeyron
Appendix E: Formulary

pv x p pref kJ kJ
ϕ=  = · ϕ1,ref = · ϕ1 cp,a = 1.004 cp,v = 1.86
dp s − s Δhv pv x + 0.622 pv p1 kg K kg K
=  =
dT v − v T (v  − v  ) kJ kJ
cliq = 4.19 cice = 2.05
kg K kg K
Joule-Thomson  
 V Ra T Rv Ra T kJ kJ
∂T δT v1+x = = 1+ ·x = · (1 + 1.608 · x) Δhv,0 = 2500 Δhm,0 = −333
=− = δh ma p Ra p kg kg
∂p h cp Steady state flow processes
v1+x = (1 + x) · v
incompressible flow
Mixture of ideal gases
Reference level
 ρ   ρ 
mi ni Vi pi p + c2 + gρz − p + c2 + gρz =
ξi = xi = σi = πi = T0,a = 273.15 K p0,a = 1 bar 2 2 2 1
mtotal ntotal Vtotal ptotal
− ρψ12 < 0
    T0,w = 273.16 K p0,w = 0.006 117 bar
ξi = 1 xi = 1 σi = 1 πi = 1
i i i i unsaturated adiabatic, frictional tube flow
πi = σi = πi  2
h1+x = ha + x · (Δhv,0 + cp,v · ϑ) 1 ṁ
h + v2 = const. = h+
1
2 A
ha = cp,a · ϑ
 Supersonic flow
ptotal = p1 + p2 + · · · + pn = pi
T p − pv
i s1+x = cp,a ln − Ra ln + 
T0,a p0,a ∂p √ √
pi Vtotal = mi Ri T ptotal Vtotal = mtotal Rtotal T   a= = κpv = κRT
T pv Δhv,0 ∂ρ s
  + xv cp,v ln − Rv ln +
Rtotal = ξi Ri Mtotal = x i Mi T0,w p0,w T0,a
c
i i Ma =
xi a
Rtotal Mtotal = RM Mtotal = Mi T p − pv Laval nozzle
ξi s1+x = cp,a ln − Ra ln +
T0,a p0,a 
  dA 1 1
T p Δhv,0 = − v dp
+ x cp,v ln − Rv ln v + A c 2 a2
  T0,w p0,w T0,a
htotal = ξi hi utotal = ξi ui
dc dA
i i fog region =− 
  c A 1 − Ma2
cp,total = ξi cp,i cv,total = ξi cv,i h1+x = ha + x · hv + (x − x ) · hliq adiabatic diffusor
i i
 Ti pi ha = cp,a · ϑ hv = Δhv,0 + cp,v · ϑ hliq = cliq · ϑ Δhs h2s − h1 c2 − c22s
Stotal = mi si (pi , T ) si − s0 = cp,i ln − Ri ln ηsD = = = 12
i
T0 p0 T Δh h2 − h1 c1 − c22
s1+x = s1+x + (x − x ) · cliq ln
T0,w
Adiabatic mixing @ ptotal = const., T = const. adiabatic nozzle
  freezing fog region
ptotal ptotal Δh h2 − h1 c2 − c22
Si,12 = m1 R1 ln + m2 R2 ln  
h1+x = ha + x · hv + (x − x ) · hice ηsN = = = 21
p1 p2 Δhs h2s − h1 c1 − c22s
965

3
966

Technical Thermodynamics

Technical combustion Energetics


Stoichiometry For ϑEG > ϑτ
mC mH2 m O2 Vk nk (ϑEG − ϑ0 ) +
ξC = =c ξH2 = =h ξO2 = =o σk = σk = xk = q =μEG · cp,EG |ϑϑEG
0
(ϑEG − ϑ0 ) + a · cAsh |ϑϑEG
0
F F F VF nF n

mS mN2 mW − HU (T0 ) − cp,F |ϑϑF0 (ϑF − ϑ0 ) − l · ξa,i · cp,i |ϑϑa0 (ϑa − ϑ0 )
ξS = =s ξN2 = =n ξW = =w i=1
F F F Omin =0.5 · xH2 + 0.5 · xCO +
mAsh 
ξAsh = =a bi zi
F + ai + − + ri · xCa Hb Oz Np Sr ,i − xO2
i
4 2 HU (T0 ) + cp,F |ϑϑF0 (ϑF − ϑ0 ) + λlmin cp,a |ϑϑa0 (ϑa − ϑ0 )
c+h+s+o+n+a+w =1 ϑad = ϑ0 +
μEG · cp,EG |ϑϑad
0
+ a · cAsh |ϑϑad
0
na,min 1
Lmin = = · Omin H0 (T0 ) = HU (T0 ) + μEG,H2 O · Δhv (T0 )
nF xO2 ,dry air
kg kg kg ma n a Ma na L
 ni
MH = 1 MC = 12 MO = 16 HUM (T0 ) = · HUM,i (T0 )
kmol kmol kmol λ= = = = nF
ma,min na,min Ma na,min Lmin i
kg kg  ni
MN = 14 MS = 32 nA = λ · na,min (1 + 1.608 · x) H0M (T0 ) = · H0M,i (T0 )
kmol kmol i
nF

mO2 ,min nEG,CO2 


= omin = 2.667 · c + 8 · h + s − o νEG,CO2 = = xCO + ai · xCa Hb Oz Np Sr ,i + xCO2 |qE | |qE + qiso |
mF nF ηK = ηF =
i HU HU
ma, min 1 
lmin = = · omin nEG,SO2 
mF ξO2 ,dry air νEG,SO2 = = ri · xCa Hb Oz Np Sr ,i ΔhEG = μEG,i · [hEG (EG ) − hEG (T0 )]
nF i i
ma l
λ= = Chemical reactions
ma,min lmin
nEG,H2 O  bi T
mA = ma + mv = ma + xma = ma (1 + x) = λma,min (1 + x) νEG,H2 O = = xH2 + · xCa Hb Oz Np Sr ,i + 
nF i
2 Hm,i (T ) = Δ0B Hm,i (T0 ) + Cp  (T − T0 )


+ 1.608 · λ · Lmin · x chemical
T0
caloric
mEG,CO2 mC
μEG,CO2 = = 3.667 · = 3.667 · c  pi  
mF mF nEG,N2
νEG,N2 = = xN2 + · xCa Hb Oz Np Sr ,i + Q̇ = ṅi Hm,i (Ti ) − ṅj Hm,j (Tj )
mEG,SO2 mS nF i
2 i,Products j,Educts
μEG,SO2 = =2· =2·s
mF mF + λ · Lmin · xN2 ,dry air
mEG,H2 O
μEG,H2 O = = 9 · h + w + λ · lmin · x  ṅi 0  ṅj
mF nEG,O2 Δ0R Hm (T0 ) = Δ Hm,i (T0 ) − Δ0 Hm,j (T0 )
νEG,O2 = = Lmin · xO2 ,dry air · (λ − 1) = Omin · (λ − 1) ṅF B ṅF B
mEG,N2 nF i,Products j,Educts
μEG,N2 = = n + λ · lmin · ξN2 ,dry air
mF = −HUM (T0 )
mEG,O2
μEG,O2 = = lmin · ξO2 ,dry air · (λ − 1) = omin · (λ − 1)
mF nEG,i νEG,i νEG,i MEG,i
 xEG,i = = = μEG,i = νEG,i ·
μEG,i = (1 − a) + λ · lmin · (1 + x) nEG νEG i νEG,i MF
i pi
Sm (T,pi ) = Sm (T,p0 ) − Rm ln .
p0
 
mEG,i μEG,i ṅEG,H2 O − xs ṅEG ṠII − ṠI = ṅi Sm,i (Ti ,pi ) − ṅj Sm,j (Tj ,pj ) .
ξEG,i = = ṅliq =
mEG i μEG,i 1 − xs i,Products j,Educts
Appendix E: Formulary

4
References

1. Hahne, E.: Technische Thermodynamik: Einführung und Anwendung. Oldenbourg, München


(2010)
2. Brown, D.: Origin. Bantam Press, London (2017)
3. Cerbe, G., Wilhelms, G.: Technische Thermodynamik: Theoretische Grundlagen und praktis-
che Anwendungen. Hanser, Berlin (2013)
4. Doering, E., Schedwill, H., Dehli, M.: Grundlagen der Technischen Thermodynamik -
Lehrbuch für Studierende der Ingenieurwissenschaften. Vieweg+Teubner, Berlin (2012)
5. Baehr, H.D., Kabelac, S.: Thermodynamik: Grundlagen und technische Anwendungen.
Springer, Wiesbaden (2012)
6. Brandt, F.: Brennstoffe und Verbrennungsrechnung. Vulkan GmbH, Essen (2000)
7. Renz, U.: Feuerungstechnik: Vorlesungsskript. Lehrstuhl für Wärmeübertragung und Klimat-
echnik, RWTH Aachen (2000)
8. Bell, I.H., Wronski, J., Quoilin, S., Lemort, V.: Pure and pseudo-pure fluid thermophysical
property evaluation and the open-source thermophysical property library coolprop. Ind. Eng.
Chem. Res. 53(6), 2498–2508 (2014). https://doi.org/10.1021/ie4033999
9. Holmgren, M.: X steam, thermodynamic properties of water and steam. https://de.mathworks.
com/matlabcentral/fileexchange/9817-x-steam--thermodynamic-properties-of-water-and-
steam (2018). Accessed 01 June 2018
10. Grothe, K.H., Feldhusen, J.: Dubbel. Taschenbuch für den Maschinenbau. Springer, Berlin
(2014)
11. Bošnjaković, F., Knoche, K.F.: Technische Thermodynamik Teil 1. Steinkopff, Darmstadt
(1988)
12. Watter, H.: Nachhaltige Energiesysteme: Grundlagen. Systemtechnik und Anwendungs-
beispiele aus der Praxis. Vieweg+Teubner, Wiesbaden (2009)
13. Freymann, R., Strobl, W., Obieglo, A.: The turbosteamer: a system introducing the principle
of cogeneration in automotive applications. MTZ Worldw. 69, 20–27 (2008)
14. Kuchling, H.: Taschenbuch der Physik. Carl Hanser GmbH & Co. KG, München (2010)
15. Schmidt, A.: Untersuchungen zum Wärmeübergang in blasenbildenden Wirbelschichten.
RWTH Aachen, Dissertation (2002)
16. Schugger, C.: Experimentelle Untersuchung des primären Strahlzerfalls bei der motorischen
Hochdruckeinspritzung. RWTH Aachen, Dissertation (2007)
17. Geller, W.: Thermodynamik für Maschinenbauer: Grundlagen für die Praxis. Springer, Berlin
(2006)
18. Stephan, P., Schaber, K., Stephan, K., Mayinger, F.: Thermodynamik - Grundlagen und technis-
che Anwendungen (Band 2: Mehrstoffsysteme und chemische Reaktionen). Springer, Berlin
(2010)
© Springer Nature Switzerland AG 2022 967
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
968 References

19. Rauschnabel, K.: Entropie und 2. Hauptsatz der Thermodynamik. http://mitarbeiter.hs-


heilbronn.de/~rauschn/5_Thermodynamik/Physik_5_7_Entropie.pdf/ (2011). Accessed 08
Feb 2011
20. Heintz, A.: Gleichgewichtsthermodynamik - Grundlagen und einfache Anwendungen.
Springer, Berlin (2011)
21. Sommerfeld, A.: Thermodynamik und Statistik, Nachdruck der 2. Auflage 1962/1977. Frank-
furt am Main, Harri Deutsch (2011)
22. Joule, J.P.: LIV. On the changes of temperature produced by the rarefaction and condensation
of air. Lond., Edinb., Dublin Philos. Mag. J. Sci. 26, Nr. 174, 369–383 (1845). https://doi.org/
10.1080/14786444508645153
23. Cerbe, G., Hoffmann, H.-J.: Einführung in die Thermodynamik. Carl Hanser, München, Wien
(2005)
24. Chemieingenieurwesen, VDI-Gesellschaft V.: VDI-Wärmeatlas. Springer, Berlin (2006)
25. Pischinger, R., Klell, M., Sams, T.: Der Fahrzeugantrieb. Bd. 1: Thermodynamik der Verbren-
nungskraftmaschine, 3. Auflage. Springer (2009). ISBN 978-3-211-99276-0
26. van der Waals, J.D.: The equation of state for gases and liquids (Reprinted from Nobel Lecture,
December 1910). J. Supercrit. Fluids 55, 403–414 (2010)
27. Redlich, O., Kwong, J.N.S.: On the thermodynamics of solutions. V. An equation of state.
fugacities of gaseous solutions. Chem. Rev. 44, Nr. 1, 233–244 (1949). https://doi.org/10.
1021/cr60137a013, PMID: 18125401
28. Peng, D.-Y., Robinson, D.: New two-constant equation of state. Ind. Eng. Chem. Fundam. 15,
02 (1976). https://doi.org/10.1021/i160057a011
29. Wisniak, J.: Daniel Berthelot. Part I. Contribution to thermodynamics. Educ. Quim. 21, 04,
155–162 (2010). https://doi.org/10.1016/S0187-893X(18)30166-6
30. MacDougall, F.: On the Dieterici equation of state. J. Am. Chem. Soc. 39(05), 1229–1235
(2002). https://doi.org/10.1021/ja02251a009
31. Onnes, K.: Heike: expression of the equation of state of gases and liquids by means of series.
In: KNAW, Proceedings, Amsterdam, vol. 4, pp. 125–147 (1902)
32. Joule, J.P., Thomson, W.: LXXVI. On the thermal effects experienced by air in rushing through
small apertures. Lond., Edinb., Dublin Philos. Mag. J. Sci. 4, Nr. 28, 481–492 (1852). https://
doi.org/10.1080/14786445208647169
33. Lemmon, E.W., McLinden, M.O., Friend, D.G.: Thermophysical properties of fluid systems.
In: NIST Chemistry WebBook, NIST Standard Reference Database, Nr. 69 (2021). https://doi.
org/10.18434/T4D303
34. Perry, R.H., Green, D.W.: Chemical Engineer’s Handbook. McGraw-Hill, New York (1997)
35. Kraus, H.: Die Atmosphäre der Erde. Vieweg, Braunschweig/Wiesbaden (2000)
36. Stierstadt, K., Fischer, G.: Thermodynamik: Von der Mikrophysik zur Makrophysik. Springer,
Berlin (2010)
37. Schröder, W.: Fluidmechanik. Mainz, Aachen (2018)
38. Ganzer, U.: Gasdynamik. Springer, Berlin (1988)
39. Müller, R.: Luftstrahltriebwerke, Grundlagen. Charakteristiken Arbeitsverhalten. Springer,
Berlin (1997)
40. Knoche, K.F.: Technische Thermodynamik. Vieweg, Braunschweig/Wiesbaden (1992)
41. Warnatz, J., Maas, U., Dibble, R.W.: Verbrennung - Physikalisch-Chemische Grundlagen,
Modellierung und Simulation, Experimente. Schadstoffentstehung. Springer, Berlin (2001)
42. Doering, E., Schedwill, H., Dehli, M.: Grundlagen der Technischen Thermodynamik.
Vieweg+Teubner, Wiesbaden (2012)
43. Joos, F.: Technische Verbrennung: Verbrennungstechnik, Verbrennungsmodellierung. Emis-
sionen. Springer, Berlin (2006)
44. Lucas, K.: Thermodynamik: Die Grundgesetze der Energie- und Stoffumwandlungen. Springer,
Berlin (2008)
45. McBride, B.J., Zehe, M.J., Gordon, S.: NASA Glenn coefficients for calculating thermody-
namic properties of individual species. In: Glenn Research Center, NASA/TP-2002-211556
(2002)
Index

A Carnot machine, 360, 367, 368, 371


Adiabatic, 43, 121 Changes of state, 83
Aggregate state, 584 adiabatic, 135, 136
Air irreversible, 94, 171, 263
humid, 565 isentropic, 267
adiabatic mixing, 610 isobar
adiabatic saturation temperature, 617 T, s-diagram, 266
balance of exergy, 601 isobaric, 86
caloric state values, 581 isochore
changes of state, 604 T, s-diagram, 266
dehumidification, 607 isochoric, 87
humidification, 614 isothermal, 85
relative humidity, 572 ploytropic, 272, 273
specific enthalpy, 583 real fluid, 275
specific entropy, 588, 592 quasi-static, 91
specific volume, 577 requirements, 93
partial pressure, 570 reversible, 94, 171, 263
saturated, 570, 573 Chemical equation, 743, 803
unsaturated, 570 Chemical reaction, 799
water content, 570, 571 Clausius-Clapeyron relation, 518
Air demand, 727, 735 Coefficient of performance, 366
Air-fuel equivalence ratio, 736 Combustion, 719
Atmosphere Components, 423
isentropic, 315 Compression, 135
standard, 837 adiabatic, 135
Avogadro’s law, 80 Compressor, 427
adiabatic, 429
Cooling machine, 43, 364, 366, 701
B Critical point, 478, 486
Bernoulli equation, 634 Cycle, 117
Boundary layer, 138 Carnot, 371, 442, 689
Boundary of a system, 3, 35, 45 clockwise, 372
counterclockwise, 375
Clausius Rankine, 445, 693
C preheating, 699
Caloric theory, 8 reheating, 696
Carnotisation, 699 clockwise, 5, 360
© Springer Nature Switzerland AG 2022 969
A. Schmidt, Technical Thermodynamics for Engineers,
https://doi.org/10.1007/978-3-030-97150-2
970 Index

counterclockwise, 5, 363, 368 reaction, 805


Joule, 443 specific, 149, 214, 225
phase change, 689 of vaporisation, 509
Seiliger, 447 total, 424, 428, 431, 436, 635
Stirling, 5, 448 Entropy, 11, 253, 330
thermodynamic, 114, 423, 442 absolute, 871–944
balancing, 281
closed system, 281
D open system, 285
Dalton’s law, 540, 571, 598 generation, 221, 348
Diabetic, 112 imperfection, 348
Diagram generation of, 11
T, s-, 262, 510 motivation, 258
T, v-, 486 of heat, 221
log p, h-, 512 process evaluation, 294
h, s-, 431, 511 specific, 217, 222
h 1+x , x-, 602, 621 Environment, 5, 96, 97
p, T -, 485 Equation of continuity, 187
p, v-, 84, 487 Equations of state, 67
p, v, T -, 480, 484 Berthelot, 498
caloric state, 510 caloric, 213, 805
Sankey, 412 combustion, 765
Differential real fluid, 222
total, 150 Dieterici, 498
Diffusor explicit, 72
adiabatic, 638 implicit, 72
Laval, 686 Peng-Robinson, 497
Dissipation, 11, 96, 111, 124, 133, 135, 401 Redlich-Kwong, 497
thermal, 75
van der Waals, 493
E Virial, 499
Efficiency, 5 Equilibrium
Carnot, 14 chemical, 59
exergetic, 408 mechanical, 56
thermal, 359, 409 thermal, 9, 28, 57
Energy, 23, 29 thermodynamic, 55, 78
changeability of, 29 Equivalent model, 145
chemical, 28 Evaporation, 569, 833
electrical, 170 Exergy, 5, 11, 142, 379
internal, 10, 29, 30 balance
kinetic, 4, 24 closed system, 401
mechanical, 23 cycle, 408
of a spring, 26, 123 open system, 404
potential, 4, 25 closed systems, 393
thermal, 27 fluid flows, 389
Energy balance, 804 heat, 381
combustion, 721, 757 loss of, 11, 398
Engine closed system, 398
combustion, 13 cycles, 405
Enthalpy of heat
absolute, 805, 808, 871–944 constant temperature, 384
extensive, 150 variable temperature, 385
formation, 805 Exhaust gas, 3
Index 971

composition, 728, 736, 745, 753 H


concentration, 731, 738 Heat, 1, 27, 119
dew point, 746 Heat exchanger, 435
Expansion, 135 balance of entropy, 438
adiabatic, 136 counterflow, 437
Joule, 215 parallel flow, 437
Expansion coefficient, isobaric thermal vol- Heat pump, 5, 16, 364, 367, 701
umetric, 154 absorption, 705
compression, 452
Heating value
higher, 771, 773, 811
F lower, 758, 766, 811, 812
Fanno correlation, 648 molar, 777
First law of thermodynamics, 10, 157, 671 Heat transfer, 8, 28, 344, 413
closed system, 159, 400, 815 Helmholtz free energy
at rest, 159 specific, 224
in motion, 164 Humid air
cycle, 407 combustion, 767
open system, 187, 403, 810
non-Steady state, 195
steady state, 203 I
Flame temperature Ideal gas, 214, 228, 473, 806
adiabatic, 782, 812 Incompressible flow, 633
Fluids Industrial revolution, 12
real, 473 Internal energy
single-Component, 473 specific, 214
Irreversibility, 5, 11
Friction, 30
Isentropic, 269
internal. see also Dissipation
Isentropic coefficient, 238
outer, 138
Fridge, 5
Fuel J
composition, 722 Joule’s paddle wheel, 29
elemental analysis, 741 Joule-Thomson coefficient, 525
fossil, 3, 719 Joule-Thomson effect, 16, 521
exergy, 837 isenthalpic throttle coefficient, 524
gaseous, 724
liquid, 723
mixture, 742 K
solid, 722 Kinetic theory, 31
Fuel-air equivalence ratio, 802
Fundamental equation of thermodynamics,
223, 264, 508 L
Liquids
incompressible, 74, 230, 244, 807
saturated, 502
G
Gas constant
specific, 79 M
Gibbs energy, 821, 822 Mach number, 680
Gibbs free energy Mass balance, 799
specific, 227 Mixture
Gibbs paradox, 555 concentration, 538
Gibb’s phase rule, 67 conversions, 548
972 Index

mass, 538 Process


molar, 539 steady state, 115, 631
pressure, 540 adiabatic, 634
volume, 539 thermodynamic, 111
gases Process value, 128, 149
adiabatic mixing temperature, 550
enthalpy, 548
entropy, 557 R
entropy generation, 555 Rankine–Hugoniot equation, 680
exergy, 557 Rayleigh correlation, 659, 669
gas constant, 546 Reference level, 31
internal energy, 548 Reservoir, 11, 357, 364
irreversibility, 552
laws of mixing, 545
molar mass, 547
S
specific heat capacity, 549
Saturated liquid line, 478
ideal gases, 537
Saturated vapour line, 478
Schwarz’s theorem, 152, 225, 227
Second law of thermodynamics, 11, 355
N
according to Clausis, 363
Nernst heat theorem, 808
according to Planck, 355
Normal shock, 668
closed system, 401, 816
No-slip condition, 95, 124
cycle, 408
Nozzle, 673
adiabatic, 639 open system, 404, 814
converging, 675 Seebeck effect, 18, 47
diverging, 681 Sign convention, 120, 125
Laval, 678, 685 Solids, 232
incompressible, 244, 807
Specific heat capacity
O averaged
Oxygen demand, 724, 732, 744 arithmetic, 240
logarithmic, 242
constant pressure, 215
P constant volume, 214
Partial energy equation, 543 molar, 871
closed system, 136, 166, 401, 403 temperature dependency, 239
open system, 206, 210 State
Partial pressure, 542 steady state, 5
ratio, 545 thermodynamic, 566
Perpetuum Mobile State of a system, 3, 35, 46
first kind, 355 State value, 9, 149
second kind, 355 caloric, 48
Phase change, 70, 476, 480 extensive, 50
Polytropic exponent, 275 intensive, 49
Potential molar, 53
chemical, 69, 541, 825 real fluids, 493
thermal, 28 specific, 51
Principle of energy conservation, 4, 23 thermal, 47
Principle of equivalence, 10, 157 Steam
Principle of mass conservation, 403, 646, saturated, 502
672, 740 Steam machine, 12
Principle of momentum conservation, 646 Steam table, 499
Proces value, 9 water, 500, 868
Index 973

Stoichiometric factor, 745 Transient balancing, 98, 112, 335


Stoichiometry, 721, 724 chemical driven, 348
Subsonic flow, 650, 661 mechanical driven, 335
Supersonic flow, 654, 664, 670 thermal driven, 342
System, 3, 35 Triple point, 485, 486
active, 45 Turbine, 423
adiabatic, 43 adiabatic, 425
closed, 42, 116 gas, 13
fully-closed, 23, 41 steam, 13
heterogeneous, 37
homogeneous, 41
multi-component, 72 V
open, 43, 412 Vaporisation, 478, 569
passive, 45 heat of, 505, 507
permeability of a, 41 isobaric, 476
single-component, 70 reversibility, 505
Velocity of sound, 645, 675, 680
Venturi nozzle, 685
T
Volume
Technical thermodynamics
specific, 47
assumptions, 60
Temperature, 5, 47
adiabatic
combustion, 782 W
combustion, 779, 782 Water
measurement of, 47 aggregate state, 567
scales, 48 anomaly, 483
Thermal engine, 5, 9, 358 Wet steam, 477, 490
Thermal machine, 512 Work, 23, 119, 122
Thermodynamic mean temperature, 289 defintion, 122
Thermodynamics effective, 130, 167
chemical, 6 mechanical, 140
equilibrium, 91 pressure, 144, 208
statistical, 6 shaft, 141
technical, 6 shifting, 142
Third law of thermodynamics, 808 technical, 124, 144
Throttle volume, 125
adiabatic, 250, 432 p, v-diagram, 128

You might also like