You are on page 1of 81

Series ISSN: 2573-3168

ZHANG
Synthesis Lectures on
Mechanical Engineering
Essential Engineering Thermodynamics
A Student’s Guide
Yumin Zhang, Southeast Missouri State University

Engineering Thermodynamics is a core course for students majoring in Mechanical and Aerospace
Engineering. Before taking this course, students usually have learned Engineering Mechanics –Statics
and Dynamics, and they are used to solving problems with calculus and differential equations.
Unfortunately, these approaches do not apply for Thermodynamics. Instead, they have to rely on
many data tables and graphs to solve problems. In addition, many concepts are hard to understand,

ESSENTIAL ENGINEERING THERMODYNAMICS


such as entropy. Therefore, most students feel very frustrated while taking this course.
The key concept in Engineering Thermodynamics is state-properties: If one knows two properties,
the state can be determined, as well as the other four properties. Unlike most textbooks, the first two
chapters of this book introduce thermodynamic properties and laws with the ideal gas model, where
equations can be engaged. In this way, students can employ their familiar approaches, and thus can
understand them much better. In order to help students understand entropy in depth, interpretation
with statistical physics is introduced. Chapters 3 and 4 discuss control-mass and control-volume
processes with general fluids, where the data tables are used to solve problems. Chapter 5 covers
a few advanced topics, which can also help students understand the concepts in thermodynamics
from a broader perspective.

ABOUT SYNTHESIS
This volume is a printed version of a work that appears in the Synthesis
Digital Library of Engineering and Computer Science. Synthesis lectures
provide concise original presentations of important research and
development topics, published quickly in digital and print formats. For
more information, visit our website: http://store.morganclaypool.com

Morgan & Claypool


store.morganclaypool.com
Essential
Engineering Thermodynamics
A Student’s Guide
Synthesis Lectures on
Mechanical Engineering
Synthesis Lectures on Mechanical Engineering series publishes 60–150 page publications
pertaining to this diverse discipline of mechanical engineering. The series presents Lectures
written for an audience of researchers, industry engineers, undergraduate and graduate
students.
Additional Synthesis series will be developed covering key areas within mechanical
engineering.
Essential Engineering Thermodynamics: A Student’s Guide
Yumin Zhang
2018

Engineering Dynamics
Cho W.S. To
2018

Solving Practical Engineering Problems in Engineering Mechanics: Dynamics


Sayavur I. Bakhtiyarov
2018

Solving Practical Engineering Mechanics Problems: Kinematics


Sayavur I. Bakhtiyarov
2018

C Programming and Numerical Analysis: An Introduction


Seiichi Nomura
2018

Mathematical Magnetohydrodynamics
Nikolas Xiros
2018

Design Engineering Journey


Ramana M. Pidaparti
2018
iii

Introduction to Kinematics and Dynamics of Machinery


Cho W. S. To
2017

Microcontroller Education: Do it Yourself, Reinvent the Wheel, Code to Learn


Dimosthenis E. Bolanakis
2017

Solving Practical Engineering Mechanics Problems: Statics


Sayavur I. Bakhtiyarov
2017

Unmanned Aircraft Design: A Review of Fundamentals


Mohammad Sadraey
2017

Introduction to Refrigeration and Air Conditioning Systems: Theory and Applications


Allan Kirkpatrick
2017

Resistance Spot Welding: Fundamentals and Applications for the Automotive Industry
Menachem Kimchi and David H. Phillips
2017

MEMS Barometers Toward Vertical Position Detecton: Background Theory, System


Prototyping, and Measurement Analysis
Dimosthenis E. Bolanakis
2017

Engineering Finite Element Analysis


Ramana M. Pidaparti
2017
Copyright © 2018 by Morgan & Claypool

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in
any form or by any means—electronic, mechanical, photocopy, recording, or any other except for brief quotations
in printed reviews, without the prior permission of the publisher.

Essential Engineering Thermodynamics: A Student’s Guide


Yumin Zhang
www.morganclaypool.com

ISBN: 9781681734231 paperback


ISBN: 9781681734248 ebook
ISBN: 9781681734255 hardcover

DOI 10.2200/S00871ED1V01Y201808MEC016

A Publication in the Morgan & Claypool Publishers series


SYNTHESIS LECTURES ON MECHANICAL ENGINEERING

Lecture #16
Series ISSN
Print 2573-3168 Electronic 2573-3176
Essential
Engineering Thermodynamics
A Student’s Guide

Yumin Zhang
Southeast Missouri State University

SYNTHESIS LECTURES ON MECHANICAL ENGINEERING #16

M
&C Morgan & cLaypool publishers
ABSTRACT
Engineering Thermodynamics is a core course for students majoring in Mechanical and Aerospace
Engineering. Before taking this course, students usually have learned Engineering Mechanics—
Statics and Dynamics, and they are used to solving problems with calculus and differential equa-
tions. Unfortunately, these approaches do not apply for Thermodynamics. Instead, they have to
rely on many data tables and graphs to solve problems. In addition, many concepts are hard
to understand, such as entropy. Therefore, most students feel very frustrated while taking this
course.
The key concept in Engineering Thermodynamics is state-properties: If one knows two prop-
erties, the state can be determined, as well as the other four properties. Unlike most textbooks,
the first two chapters of this book introduce thermodynamic properties and laws with the ideal
gas model, where equations can be engaged. In this way, students can employ their familiar
approaches, and thus can understand them much better. In order to help students understand
entropy in depth, interpretation with statistical physics is introduced. Chapters 3 and 4 discuss
control-mass and control-volume processes with general fluids, where the data tables are used to
solve problems. Chapter 5 covers a few advanced topics, which can also help students understand
the concepts in thermodynamics from a broader perspective.

KEYWORDS
thermodynamics properties, thermodynamics laws, control mass processes, control
volume processes, information, equilibrium, chemical potential, dissipative struc-
tures
vii

Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

1 Thermodynamics Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Criteria of Ideal Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Ideal Gas Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Pressure and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Temperature and Zeroth Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Internal Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Thermodynamics Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1 First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Second Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Third Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Maxwell Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3 Control Mass Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25


3.1 State and Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Phase Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Isochoric Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4 Isobaric Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.5 Isothermal Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.6 Adiabatic and Isentropic Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4 Control Volume Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37


4.1 Mass Rate Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 Energy Rate Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Nozzle and Diffuser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
viii
4.4 Compressor and Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.5 Throttle Valve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.6 Entropy Rate Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5 Advanced Topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.1 Entropy and Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.2 Equilibrium in Isolated Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.3 Equilibrium with Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.4 Crystal Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.5 Chemical Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.6 Dissipative Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.7 Socioeconomic Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

Author’s Biography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
ix

Preface
Thermodynamics is a notoriously challenging course, and anyone exposed to it can testify to
this fact. Before encountering Thermodynamics, most engineering students have learned En-
gineering Mechanics—Statics and Dynamics. These are very elegant courses, and most of the
problems can be solved by applying Newton’s laws with the tools of calculus. Thermodynamics
is very different, and it relies on look-up tables and graphs to solve most of the engineering
problems. Therefore, students often feel at a loss in this course, and some of them even lose in-
terest in engineering altogether. When I was an undergraduate student majoring in Engineering
Mechanics, I also hated this course. Actually, I was so disappointed that I decided to switch to
Physics.
Fate gave me a second chance to encounter Thermodynamics. In the Fall semester of
2008 I received the teaching assignment of this course at Southeast Missouri State University.
With two decades of research experience, my perspective had changed significantly: Analytical
solution is not the norm but the exception, and almost all the practical problems in science and
engineering have to rely on numerical solutions. When I relearned this topic in preparing my
lectures, I realized that I had missed something important when I was a young student. Actually,
Thermodynamics is a beautiful course, and it can be learned easily. At the end of the Fall 2016
semester, I asked my students in this course to compare the difficulty level of Thermodynamics
with that of Engineering Mechanics—Dynamics, and the survey result is shown in the following
diagram.
5
Number of Students

Dynamics
4
Thermodynamics
3

0
1 2 3 4 5
Difficulty Level

The numbers on the horizontal axis indicate the difficulty level: “1” stands for the least
challenging and “5” for the most challenging. The numbers on the vertical axis refer to the
number of students who selected a specific challenge level. Twelve students took the survey, but
two of them had not learned Dynamics yet, so they are excluded. The average difficulty level
for Dynamics was 3.5, and the average for Thermodynamics was 3.3. It indicates that the latter
x PREFACE
is easier than the former. Furthermore, none of the students selected 5 (most challenging) for
Thermodynamics.
Inspired by this survey result, I would like to share my approach in helping students learn
Thermodynamics to a wider audience. I hope that the students struggling with this course can
be relieved from the hardship. Furthermore, I hope that they can also appreciate the beauty of
Thermodynamics and develop a deeper understanding of nature and human society.
The first two chapters introduce thermodynamic properties and laws, where the discussion
is limited to ideal gases. In this way, equations can be derived and also be applied in solving
problems. This is a familiar approach for most students, and they can understand these basic
concepts more easily. However, the working substances in most engineering applications cannot
be analyzed with the ideal gas model, but the overarching concepts are relevant.
The central idea is the state-properties relationship: Each state corresponds to a group
of six properties—just like six family members. Based on this understanding, we then move to
practical applications. Chapter 3 discusses control-mass processes, where the fluid is confined in
sealed containers, such as the piston-cylinder system in internal combustion engine. Chapter 4
discusses control-volume processes, where the fluid flows into and out of the devices, such as a jet
engine. In Chapter 5, Thermodynamics is applied to a wider range of fields, such as chemistry,
biology, economy, sociology and information theory.
What is Thermodynamics? It depends on to whom you are asking this question. Here are
some of the answers.
• Mathematician: A zoo of differential relationships.
• Physicist: Energy and entropy.
• Engineer: Heat and work.
• Philosopher: Everything!

Yumin Zhang
September 2018
xi

Acknowledgments
I would like to thank all the students who took this class with me, since they gave me
the opportunity to discover the beauty of Thermodynamics. Special appreciation goes to two
of them: Robert Roach and Varun Sadaphal—each of them did a thorough job in editing the
manuscript in two iterations. In addition, I also appreciate the support from my wife, who en-
dured many lonely hours while I was writing this book.

Yumin Zhang
September 2018
1

CHAPTER 1

Thermodynamics Properties
When one describes the motion of a particle, the parameters used include velocity, momen-
tum, kinetic energy, etc. Engineering thermodynamics mainly deals with fluid (liquid and gas),
which consists of a large number of molecules in random motion. Therefore, we need a new
set of parameters to describe its state, which are called the properties of the fluid: pressure (P ),
temperature (T ), specific volume (v ), entropy (s ), internal energy (u), and enthalpy (h). In this
chapter, we will introduce five of the properties and leave out entropy for next chapter. With
the ideal gas model, many relationships between these properties can be expressed with simple
equations.

1.1 CRITERIA OF IDEAL GAS


Any substance in the gas form can be investigated by using the ideal gas model, provided some
criteria are met. At the microscopic level, there is only one criterion: the distance between the
molecules is large enough so that their interactions are negligible, except for the collisions. Fig-
ure 1.1 shows the generic potential energy between two molecules. When two molecules get too
close (r < ro ), there is a repulsive force between them. On the other hand, when the distance of
separation becomes large (r > ro ), an attraction force shows up. In addition, the force between
the molecules can be found from the potential energy: FE D rV .Er /. In the isotropic case, the
gradient operator becomes a simple derivative. When the distance between molecules gets very
large, the potential energy curve is almost flat and the interaction force approaches zero. This is
the situation of ideal gas.
Because the average distance between molecules cannot be measured easily, criteria from
the macroscopic measurable parameters are needed. The property of gas directly related to the
inter-molecular distance is the density (), which is defined as the mass of the substance per unit
volume. However, in thermodynamics, the inverse of density, which is called specific volume (v),
is adopted as a property of substance. It is the volume of a substance per unit mass. If the average
distance between molecules is very large, each molecule occupies a rather large volume, so the
density will be very low, and the specific volume should be very high. In the SI unit system,
the unit of specific volume is m3 /kg. For example, air at 25ı C and 1 bar (close to atmosphere
pressure) has density of 1.169 kg/m3 , and its specific volume is 0.8554 m3 /kg.
Another property of gas is pressure (P ), which is closely related to the specific volume or
density. For example, one can compress air into a bicycle tire by pressing down the piston of an
air pump. It seems there is an inverse relationship between pressure and specific volume, since
2 1. THERMODYNAMICS PROPERTIES

ro
0 r

Eo

Figure 1.1: Potential energy vs. intermolecular distance.

the pressure increases when its volume is reduced. Because the specific volume of ideal gas is
required to be high, the criterion on pressure should be low.
Air at atmospheric pressure is an example of ideal gas, as well as all its components in this
condition, such as nitrogen (N2 ) and oxygen (O2 ). However, we don’t think that its density or
pressure is very low. In this case, we might look at another property of gas: the temperature. The
boiling points of N2 and O2 are 77 K and 90 K, respectively. From this example we can reach
another conclusion: the temperature of ideal gas is usually much higher than its boiling point.

1.2 IDEAL GAS LAW


In the previous section, we discussed the criteria of ideal gas from the point of view of macro-
scopic properties. However, one factor was missing, the substance, which is the missing link
between specific volume and intermolecular distance. For example, with 1 kg of gas contained in
a tank with the volume of 1 m3 , its specific volume is 1 m3 /kg, regardless of the molecular mass.
However, the average molecular distance will depend on the substance. For example, the molec-
ular masses of hydrogen and water are approximately 2 g/mol and 18 g/mol, which indicates that
the number of hydrogen molecules per kg is nine times the number of water molecules with the
same mass. In this situation, the volume each hydrogen molecule occupies is only one-ninth of
the p
volume of each water molecule. Therefore, the ratio of average intermolecular distances is
1 W 3 9.
1.2. IDEAL GAS LAW 3
If we take all these factors into consideration, a parameter called compressibility can be
defined:
Pv PV
ZD D : (1.1)
RT mRT
In this equation, the parameter R is different from the universal gas constant Ro D
8:3145 J/mol  K; instead, it is defined as R D Ro =M , where M is the molecular mass in g/mol.
In engineering thermodynamics larger units are preferred, so the unit of R is in kJ/kg  K. Be-
cause molecular mass varies for different substances, R is no longer a universal constant. For
example, R D 0:2968 kJ/kg  K for nitrogen and R D 0:2598 kJ/kg  K for oxygen.
Figure 1.2 shows the compressibility of nitrogen at different temperature and pressure. As
pressure reduces, the compressibility approaches unity. On the other hand, as the temperature
rises, the range of pressure with unity compressibility increases. For example, at room tempera-
ture (T D 300 K), the compressibility is unity up to 8 MPa. If the temperature drops to 150 K,
the range of the pressure with unity compressibility is below 1 MPa. The shaded area in the
diagram is in liquid phase.

T1 < T2 < T3 < T4 < T5


2
Compressibility, Pv/RT

T1 T5
1
T4

Saturated Vapor T3
T2
Saturated Liquid
0
Critical Point
Pressure

Figure 1.2: Compressibility of nitrogen vs. pressure at different temperatures.

For ideal gases, Z D 1 is assumed, and then the ideal gas law can be derived:
P V D mRT or P v D RT: (1.2)
When one property changes, the other properties will change at the same time. Such a transition
is called a process. In a control mass process, where the substance is confined in a sealed container
and thus the mass remains the same, the following relationship exists:
P1 V1 P2 V2
D D mR: (1.3)
T1 T2
4 1. THERMODYNAMICS PROPERTIES
If one of the three properties remain the same, simpler equations can be derived:
• Isothermal Process (T1 D T2 ): P1 V1 D P2 V2
V1 V2
• Isobaric Process (P1 D P2 ): D
T1 T2
P1 P2
• Isochoric Process (V1 D V2 ): D .
T1 T2
These processes can be illustrated in a P –V diagram, and an example is shown in Fig. 1.3.
In the isothermal process (1 ! 2), P V D mRT is a constant, so the pressure is inversely pro-
portional to the volume. In the isobaric process (2 ! 3), the pressure remains constant. In the
same way, the volume remains constant in the isochoric process (3 ! 1).

P 1

3 2

Figure 1.3: Processes in P –V diagram.

1.3 PRESSURE AND WORK


From a macroscopic point of view, pressure can be understood as the ratio between a force and
the action area: P D F=A. For example, the tip of a nail is very sharp, but its head is rather
broad. When a hammer hits the head, the pressure there is not very high so that no damage is
caused. However, with the same amount of force transferred to the tip, the pressure becomes
much higher due to a decrease of the action area, so it can penetrate into wood boards easily.
When a container is filled with gas, pressure can build up. An everyday life example is the
tire pressure of automobiles, and it is a good idea to check it regularly, especially before setting
off for a long trip. The SI unit of pressure is N/m2 or Pa, in memory of French scientist Blaise
Pascal who made significant contributions in hydrodynamics in the 17th century. However, this
unit is too small, and another unit widely used in engineering is bar (105 Pa), which is very close
1.3. PRESSURE AND WORK 5
to the atmosphere pressure at sea level (1.013 bar). The English unit of pressure is psi (pound
per square inch), which is about 0.069 bar. The standard atmosphere pressure is approximately
14.7 psi.
Gas molecules can be viewed as small ping-pong balls, which strike the walls of the con-
tainer constantly, and gives rise to the force and pressure. Assume the momentum of each
molecule is mv and the duration of collision is  . The force of elastic collision at the wall can
be found as f D 2mv cos = , where  is the angle between the velocity and normal direction.
Because the number of molecules is so huge, the measured pressure is the collective effect of
such collisions.
In an isothermal compression process, the density of molecules increases, so the number
of collisions gets higher, which causes an increase of pressure. On the other hand, when tem-
perature rises in an isochoric process, the average velocity of the gas molecules increases, so the
striking force f gets larger, which can also cause the rise of the pressure.
If a part of the container wall is movable, such as the piston-cylinder system shown in
Fig. 1.4, work can be done by the gas inside the system. Assume the cross-section area of the
piston is A, and the displacement is x , the work done can be found as:
Z 2 Z 2
ıW D F  x D P  A  x D P  V ! W1!2 D P dV D m P dv: (1.4)
1 1

• Isobaric process: W1!2 D P .V2 V1 / D mR .T2 T1 /


• Isothermal process: W1!2 D mRT ln .V2 =V1 /.

Figure 1.4: Piston-cylinder system.

Historically, thermodynamics was developed in the background of steam engines. There-


fore, the work is considered positive when the piston is pushed outward. In a P –V diagram,
the work can be visualized as the area under the curve, but it could be negative depending on
the direction of motion in the process. In the isothermal process 1 ! 2 shown in Fig. 1.3, the
volume increases and thus the work is positive, and it is represented as the area between this
curve and the horizontal axis. In the isobaric process 2 ! 3, the volume decreases and thus the
work is negative, but its absolute value is still equal to the area of the rectangle formed between
the line and the horizontal axis. No work is done in an isochoric process, since there is no change
in volume.
6 1. THERMODYNAMICS PROPERTIES
1.4 TEMPERATURE AND ZEROTH LAW
All of us are very familiar with the concept of temperature, and it is heard every day in the
weather forecast. However, the physics of temperature is rather convoluted. Unlike pressure,
which is a useful concept for fluid, temperature can be defined for any states of a substance.
Even vacuum has a temperature. For example, the background of the universe has a temperature
of 2.725 K. It is rather amazing that a parameter can be defined in so many ways and yet remains
consistent in describing these phenomena.
In solid materials, the atoms are vibrating around their equilibrium positions, and the
temperature describes the intensity of such vibrations. There are two classes of good thermal
conductors: the first are the ones which have very high stiffness, such as diamond, where the
vibration energy can be transferred through lattice vibration very effectively. The second is good
electric conductors, where electrons serve as the medium in transferring the thermal energy. In
fluids, the situation is quite different, where molecules can move about, as well as rotate and
vibrate internally. In such a situation, temperature can represent its kinetic energy.
In quantum systems, temperature is associated with the distribution of particles at dif-
ferent energy states. Figure 1.5 shows the conduction band and valence band of intrinsic semi-
conductors. At T D 0 K, all the electrons are in the valence band, and the conduction band is
empty. However, as temperature rises, some electrons can obtain enough energy and jump up to
the conduction band, and the conductivity will increase significantly. This effect can be utilized
in bolometers for detection of infrared radiation.

Eg

Figure 1.5: Conduction band and valence band of intrinsic semiconductors.

When two objects A and B with different temperatures are put into contact, energy will
transfer from the object with a higher temperature to the one with a lower temperature, and
eventually equilibrium is achieved with a shared temperature. Now these two objects are sep-
arated, and the object A is put into contact with another object C and they are happened to
be in equilibrium. In this case, it can be asserted that the objects B and C should also be in
equilibrium. This is the zeroth law of thermodynamics.
1.5. INTERNAL ENERGY 7
Although the zeroth law seems trivial, it provides the foundation to define temperature
across different systems. For example, the object A is in solid phase, the object B is in gas phase,
and object C is in plasma phase. With the temperature defined, we can predict how the energy
will be transferred when any two objects are put into contact. If energy is imagined as water
stored in a container, temperature is equivalent to the water level. If the containers B and C are
compared with container A separately, and it is found that both B and C have the same water
level with A (TB D TA , and TC D TA ), and then it is certain that the containers B and C also
have the same water level (TB D TC ). Based on this law, thermometers can be implemented for
temperature measurement.

1.5 INTERNAL ENERGY


When heat is absorbed by gas in a sealed container, the energy received is stored in the form
of internal energy, which is the microscopic kinetic and potential energy of the molecules. This
relationship can be expressed in the following formula:

Q1!2 D U D U2 U1 : (1.5)

For ideal gas, the average distance between molecules is rather large and thus the intermolecular
potential energy can be ignored, so the internal energy is just the kinetic energy at the molecular
level. For this, there are three modes of motion: translation, rotation, and vibration.
Inert gases are special, since each molecule has only one atom, no rotation and vibration
is involved. In this case, its internal energy is only expressed in the kinetic energy in translation.
From statistic mechanics, each degree of freedom of a molecule has the energy of 12 kB T , so the
internal energy of an inert gas molecule is 32 kB T . The factor 3 in this formula represents the
three dimensions in space, and kB D 1:38  10 23 J/K is the Boltzmann’s constant. This simple
relationship can be expressed as U D 23 N kB T , where N stands for the total number of atoms.
Thermodynamic properties are classified into two categories: intensive and extensive. The
extensive properties are proportional to the mass. For example, the internal energy will double if
the mass is doubled, provided the temperature remains the same. On the other hand, the inten-
sive properties are independent of mass. For example, one can measure the water temperature of
a swimming pool by taking a small sample of the water for measurement. Both temperature and
pressure are pure intensive properties that can be measured easily, and thus they are adopted as
the two parameters for many thermodynamics tables.
In addition, extensive properties can be converted into intensive properties. In this case,
they are represented by upper- and lowercases letters, respectively. For example, we can convert
the extensive internal energy into intensive internal energy: u D U=m D 32 N k T . In this way,
m B
these intensive parameters can be tabulated and compared easily.
The equation of the specific internal energy can be simplified: for N D NA D 6:022  1023
(1 mol), the corresponding mass is the molecular mass m D M , and then NA kB D Ro . Therefore,
8 1. THERMODYNAMICS PROPERTIES
the intensive internal energy of inert gases can be found:

3 NA kB 3 Ro 3
uD T D T D RT: (1.6)
2 M 2M 2

A related parameter can be defined, and it is called specified heat capacity (usually called specific
heat) at constant volume (Cv ), which is the amount of heat absorbed for 1 kg of fluid sealed in a
rigid container when the temperature rises one degree. This definition can be expressed below:
 
du
Cv D : (1.7)
dT v

For inert gases, the specific heat is a constant: Cv D 32 R. Keep in mind, R D Ro =M is a substance
dependent constant. The emphasis is that it does not change with temperature, which is unique
among the various gases.
From a certain point of view, the concept of specific heat is similar to capacitance used in
electric circuits. As we know, a capacitor can be imagined as a container of electric charges. In
the same way, an ideal gas can also be modeled as a container of heat. For inert gases, the shape
of such a container is very simple, just like a graduated cylinder, which is shown in Fig. 1.6a.
If the intensive internal energy is represented as the amount of water in the cylinder and the
temperature is equivalent to the water level, the specific heat is the cross-section area, which is
independent of temperature. In this case, the change of internal energy can be calculated very
easily: U D mCv T . However, for other types of ideal gases, the situation is more compli-
cated, and it is more like the situation of a wine glass shown in Fig. 1.6b.

(a) (b)

Figure 1.6: Specific heat capacity vs. temperature.


1.6. ENTHALPY 9
In general, the specific heat (cross section area) is a function of temperature (height), so
the change of internal energy must be calculated with an integral:
Z T2
U D m Cv .T /d T: (1.8)
T1

For gases of complex molecules, the modes of motion are more complicated, because the ro-
tation and vibration need to be taken into account. The energy of intra-molecular vibration is
quantized, which can be described by a series of energy levels. When the temperature is low
.kT  „!0 /, the energy is at the lowest level, which makes no contribution to the internal en-
ergy and specific heat. In this case, we can still use the graduated cylinder model in Fig. 1.6a to
describe the internal energy of ideal gases.
For gases of diatomic molecules (CO and N2 ) or any straight chain molecules (CO2 ), there
are only two degrees of freedom of rotation, so the formula of the internal energy and specific
heat are u D 25 RT and Cv D 52 R. For example, let the carbon dioxide molecule (C=O=C) align
along the z -axis, then the rotational kinetic energy is associated with the rotations along x - and
y -axes, because the rotation along the z -axis has no contribution to the kinetic energy (zero
moment of inertia). For more complicated molecules, such as H2 O, there are three degrees of
freedom in rotation, so its formula of internal energy and specific heat at low temperature are
u D 3RT and Cv D 3R.
When temperature rises to a certain point .kT  „!0 /, the contribution from the vi-
bration modes will kick in. In this way, the heat energy absorbed can be stored in the form of
molecular vibration, and thus the specific heat capacity becomes higher. Therefore, the internal
energy becomes a nonlinear function of temperature, so the situation is more like the wine glass.
In the same way, the specific heat also becomes temperature dependent.

1.6 ENTHALPY
The concept of internal energy can be directly applied in the case without volume change, but
it is not very convenient in other situations. For example, consider solid and liquid substances.
It is unrealistic to require the volume remaining the same when changing the temperature. On
the other hand, the pressure can be assumed unchanged in these situations. Therefore, the heat
capacity defined in an isobaric process is more useful. However, when the substance expands
or contacts, work is involved. In other words, some of the heat absorbed will be converted to
work, and the remaining part is stored in the form of internal energy. Now a new property can
be defined: H D Q D U C PV . The property H in this formula is called enthalpy, which
is closely related to the internal energy:

H D U C PV or h D u C P v D u C RT: (1.9)
10 1. THERMODYNAMICS PROPERTIES
Specific heat capacity in an isobaric process can be defined in a similar way:
 
dh
Cp D D Cv C R: (1.10)
dT p

At low temperature, this specific heat capacity with constant pressure also has very simple ex-
pression:
• Mono-atomic molecules: Cp D 52 R

• Straight molecules: Cp D 72 R

• Complex molecules: Cp D 4R .

Table 1.1 shows the ratios of Cp over R of a number of gasses at the condition of 25ı C and
100 kPa. For simple molecules, the data agree with the theory quite well. However, for complex
molecules, the vibration effect kicks in at a lower temperature, which causes an increase of the
specific heat capacity.

Table 1.1: Ratios of Cp over R of a number gasses at the condition of 25ı C and 100 kPa

Gas He Ne H2 O2 CO2 H2O C2H4


CP (kJ/kg K) 5.193 1.030 14.209 0.922 0.842 1.872 1.548
R (kJ/kg K) 2.0771 0.4120 4.1243 0.2598 0.1889 0.4615 0.2964
CP /R 2.500 2.500 3.445 3.549 4.457 4.056 5.223

Figure 1.7 shows the specific heat capacities of various gases as a function of temper-
ature. First, we notice that the inert gases have constant value and they are exactly at 2:5R,
since there is no rotation and vibration. For all the other kinds of gases, the specific heat in-
creases with temperature. Furthermore, the gases with more complex molecular structures have
stronger dependence on temperature, since more vibration modes may be activated. However,
at low temperatures, their values are determined by the degrees of freedom in the translation
and rotation modes.
For ideal gases, the internal energy and enthalpy are functions of temperature only: u D
u.T / and h D h.T /. In later chapters we will discuss non-ideal gases, where the inter-molecular
potential energy must be taken into account. Since it is a function of the inter-molecular distance,
as shown in Fig. 1.1, the internal energy and enthalpy will also change with specific volume
or pressure. Because pressure is much easier to measure than specific volume, for non-ideal
gases these two properties are generally tabulated as functions of both temperature and pressure:
u D u.T; P / and h D h.T; P /.
1.6. ENTHALPY 11

8
C2H4
7
CO2

H2O
C p0
5
R
O2

4
H2

3
Ar, He, Ne, Kr, Xe

2
0 500 1000 1500 2000 2500 3000 3500

T [K]

Figure 1.7: Specific heat capacity for some gases as a function of temperature.
13

CHAPTER 2

Thermodynamics Laws
In classical mechanics, Newton’s three laws on motion play the central role. Similarly, there are
four laws in thermodynamics. Since the zeroth law was already covered in the previous chapter,
we will discuss the remaining three laws in this chapter. The first law is on energy conservation,
which can be expressed in an equation. On the other hand, the second and third laws tell us the
limits in thermodynamics.

2.1 FIRST LAW OF THERMODYNAMICS


Simply speaking, the first law of thermodynamics is just the principle of energy conservation.
We have learned its special forms in different circumstances. For example, if the work of non-
conservative forces can be ignored, the mechanical energy—summation of kinetic and potential
energy—is conserved. In dealing with heat transfer problems, we also assume that the total
amount of heat energy is conserved. In thermodynamics, the boundary between mechanics and
heat is broken, so both of them are included in a general form:
Q1!2 D U C W1!2 : (2.1)
As shown in the above equation, there are three variables in the first law of thermodynamics,
so you can find one of them if the other two can be figured out. If a heat engine is imagined
as a human body, the heat absorbed .Q1!2 / is equivalent to the energy that one intakes from
the food, and the work .W1!2 / is the energy consumed in physical activity. If the metabolic
processes of human body are ignored, the change of internal energy .U / is similar to the fat
accumulated or reduced. In other words, if one eats less and works out a lot, he will lose weight,
and vice versa.
In this equation the heat transfer (Q1!2 ) and work (W1!2 ) are process dependent. For
example, in Fig. 2.1 the work W1!2 can be calculated in two different ways.
• Path A—directly following the process (1 ! 2).
• Path B —following an isochoric process (1 ! 3) first and then the isobaric process (3 ! 2).
As we know, the amount of work is equal to the area underneath the curve. Therefore, the
A B
work done in path A is greater than the work done in path B , W1!2 > W1!2 .
On the other hand, the specific internal energy (u) is a state property, which is independent
of the path: U D m.u2 u1 /. In other words, regardless of the process, the change of internal
energy is always equal to the difference between the final state and the initial state.
14 2. THERMODYNAMICS LAWS

P
1

3 2

Figure 2.1: Work in P –V diagram.

For ideal gases, we can use a polytropic process to represent many different processes, and
a general relationship can be expressed as:

P  V n D constant (2.2)

• n D 0: P D constant (isobaric process)


• n D 1: T D constant (isothermal process)
• n D k: P V k D constant (adiabatic process, k D Cp =Cv )
• n D 1: V D constant (isochoric process).
The relationship of adiabatic process will be discussed in next section. Plugging this gen-
eral expression (Eq. (2.2)) into the formula of work (Eq. (1.4)), the following equation can be
derived:
P2 V2 P1 V1 mR.T2 T1 /
W1!2 D D .n ¤ 1/: (2.3)
1 n 1 n
For isothermal process (n D 1), the result is:
   
V2 P1
W1!2 D mRT ln D mRT ln .n D 1/: (2.4)
V1 P2

2.2 ENTROPY
The first law of thermodynamics indicates that the quantity of the energy should be conserved
when energy is transformed between different forms. For example, the electric energy consumed
2.2. ENTROPY 15
in a heater is equal to the same amount of heat it gives off. However, in a power plant the amount
of electric power generated is less than the amount of power consumed from burning fossil fuels.
In this process, there is a limit of efficiency in physics that cannot be overcome by improvements
and innovations in technology. Besides the aspect of quantity, energy also has quality. When a
high-quality energy form (electric) is transformed into low-quality energy form (thermal), the
efficiency can approach 100%. However, it is not true for the reverse process.
Heat is the energy form with the lowest quality, but subtle difference within this category
can be further distinguished, such as the temperature of the heat source. For example, two objects
with different temperatures are put into contact, and heat will be transferred from the one with
the higher temperature to the one with the lower temperature. One might ask: “Is it possible
to reverse the direction of heat transfer?” If that happens, the first law of thermodynamics still
holds, since the total amount of energy is not changed.
In order to set up rules in dealing with these phenomena, a new concept is needed. Entropy
was introduced in the 19th century, when steam engine played the dominant role in the industrial
revolution. At that time, the microscopic theory was not available, so people could only define
it in the incremental format:
ıQ
dS D : (2.5)
T

This equation indicates that the change in entropy (dS ) is equal to the heat transferred .ıQ/
divided by the temperature of the object (T in Kelvin). Since the temperature is likely to change
with the heat transfer process, it is valid only when the amount of heat is very small, and the
process is reversible.
With the concept of entropy defined, we can revisit the heat transfer process between two
objects. Assuming the temperatures of the two objects are TA and TB .TA > TB /. In the normal
heat transfer process, the entropy change of the system with both objects is:

ıQ ıQ
dS D dSA C dSB D C > 0: (2.6)
TA TB

In this case, the object A with a higher temperature gives off a certain amount of heat to the
object B with a lower temperature, which is reflected in the signs of ıQ in the equation. If
the reverse process happens, the signs should be flipped, and the change of entropy is negative.
Therefore, we can distinguish between these two processes from the sign of entropy change.
Phase change is a very interesting phenomena, where the temperature remains constant.
In this case, the change in entropy can be calculated very easily: S D Q=T . For example,
when water evaporates into steam, it needs to absorb a lot of heat (Q > 0). As a result, the
entropy increases significantly. In the reverse process, steam releases a lot of heat (Q < 0) when
it condenses into water, so the entropy decreases.
16 2. THERMODYNAMICS LAWS
With a simple transformation, we can find a way to calculate the amount of heat transfer
from entropy change in reversible processes:
Z 2 Z 2
ıQ D T dS ! Q1!2 D T dS D m T ds: (2.7)
1 1

The counterpart to this formula is Eq. (1.4), where work can be calculated in a similar way.
Therefore, P –V diagram can be used in calculating work, and T –S diagram can be used in
calculating heat. Just like T , P , and v , the intensive form of entropy (s ) is also a state property.
With the addition of u and h, these six properties are tabulated for engineering applications.
Beginning next chapter, we will rely on these data to solve problems.
Equation (2.7) is valid only in reversible processes, which is a good approximation in most
practical processes. In a process without any heat transfer, which is called adiabatic process, the
entropy should remain constant, provided it is a reversible process. Such a process is also called
isentropic, which is represented in a vertical line in T –S diagram. Unlike internal energy and
enthalpy, entropy also depends on volume, even for ideal gas.
An interesting example of an irreversible process is the adiabatic unstrained expansion
process. An insulated container is separated into two chambers separated by a membrane. Ini-
tially, gas is contained in one chamber, and the other one is in vacuum. If a hole is poked at
the membrane, the gas can occupy both chambers. In this irreversible process, the entropy is
increased, although there is no heat transfer. Therefore, the adiabatic process is not identical to
the isentropic process; they are equivalent only in reversible processes.
For ideal gas, we can derive a formula to calculate entropy change. First, we start with the
first law in the differential format:
ıQ D d U C ıW ! mT ds D mCv d T C mP dv: (2.8)
By applying the ideal gas law, P v D RT , the change of entropy can be found:
dT dv
ds D Cv CR : (2.9)
T v
At low temperature or within a small temperature range, the specific heat Cv is a constant, so a
simple expression can be found:
T2 v2 T2 V2
s D Cv ln C R ln D Cv ln C R ln : (2.10)
T1 v1 T1 V1
In the process of adiabatic unstrained expansion process, the temperature of ideal gas is
unchanged. Following the first law, mu D Q1!2 W1!2 D 0, since no work is done. For ideal
gas, internal energy depends on temperature only, so there is no temperature change. Therefore,
we can find the change of entropy in this process easily: S D ms D mR ln.V2 =V1 /.
If an adiabatic expansion process is reversible, such as in a piston-cylinder system, the
result will be very different. Because work is done to the environment with the cost of reduction
2.3. SECOND LAW OF THERMODYNAMICS 17
of internal energy, the temperature of ideal gas is lowered. In such an isentropic process, the
change of entropy vanishes following Eq. (2.5). In this case, we can find the relationship between
P and V by plugging this relationship T2 =T1 D .P2 =P1 /  .V2 =V1 / into Eq. (2.10):

P2 V2
s D Cv ln C Cp ln D 0 ! P1 V1k D P2 V2k : (2.11)
P1 V1

In this equation, k D Cp =Cv . Many people assume that Eq. (2.11) is valid as long as it is an
ideal gas. However, the derivation process shows that it cannot be applied in the situation when
the specific heat capacity varies significantly in the process. As we discussed in Section 1.5, only
inert gases have constant specific heat capacities. However, if the temperature change is small,
one can assume that the change in specific heat is insignificant.

2.3 SECOND LAW OF THERMODYNAMICS


Originally, the second law of thermodynamics was stated in empiric statements. The two most
important applications in this area are heat engines and heat pumps, which include air condi-
tioning and refrigerators. The principle diagrams of heat engine and heat pump are shown in
Fig. 2.2. The Kelvin–Planck statement is on heat engines: “It is impossible to absorb heat from
one heat source and convert all of it into work.” In other words, the operation of a heat en-
gine needs at least two heat sources—it absorbs heat from a high temperature source and rejects
heat to a low temperature source. From another point of view, this statement indicates that the
efficiency of heat engine cannot reach 100%. The Clausius statement is on heat pumps: “It is
impossible to pump heat from a cooler body to a hotter body without work.”

TH TH

QH QH

W W

QL QL

TL TL

Figure 2.2: Principle diagrams of (a) heat engine and (b) heat pump.

After the concept of entropy is introduced, the second law can be stated in a more elegant
way: “The entropy in an isolated system cannot decrease spontaneously.” Entropy increase is
always associated with the degradation of energy, so the most efficient system will keep entropy
18 2. THERMODYNAMICS LAWS
T

a b
TH

TL d c

S
S1 S2

Figure 2.3: A Carnot cycle for heat engine in T –S diagram.

unchanged in a cycle. A Carnot cycle can achieve this, so it is the most efficient system. A Carnot
cycle includes two isothermal processes and two isentropic processes, which is shown in Fig. 2.3.
A Carnot cycle can be in two different orientations: the clockwise one for heat engines,
and the counter-clockwise one for heat pumps. In the cycle shown in Fig. 2.2a, the heat engine
absorbs heat from a large heat source at a high temperature TH , and rejects heat to a large
low temperature heat source TL . If the capacity of the heat sources is not large enough, the heat
exchange will cause a temperature change, and then the two processes with heat exchange are no
longer isothermal processes. In the two isentropic processes, as we discussed in previous section,
no heat exchange happens.
The amount of heat exchange can be calculated with Eq. (2.7): jQH j D TH .S2 S1 / and
jQL j D TL .S2 S1 /. With the first law of thermodynamics, the amount of work can be found:
jW j D jQH j jQL j D .TH TL /.S2 S1 /, which is equal to the area of the rectangle in the
T –S diagram. It can be verified easily that the entropy change is equal to zero in a complete
cycle:
QH QL
S D C D 0: (2.12)
TH TL
The efficiency of a cycle is defined as the ratio of desired outcome over the input/cost, so different
definitions are used:
jW j jQL j TL
• Heat Engine: D D1 D1
jQH j jQH j TH
jQL j jQL j TL
• AC/Refrigerator: ˇD D D
jW j jQH j jQL j TH TL
jQH j jQH j TH
• Heat Pump: D D D .
jW j jQH j jQL j TH TL
For heat engines, the cost is the heat derived from burning fuel, so the denominator is QH .
The formula indicates that the efficiency cannot reach 100%, unless we could access heat sources
2.4. THIRD LAW OF THERMODYNAMICS 19
with TL D 0 K or TH D 1 K. This conclusion is consistent with the Kelvin–Plank statement of
the second law. For an AC and a refrigerator, the desired outcome is the heat pumped out of the
low temperature body, so the numerator is QL . The narrowly defined heat pump is a device used
for heating buildings in winter, which can be considered as an AC installed in the backward way.
Actually, it can suck heat out of the cooler air outside of the building and pump it inside. It is
more efficient than using electric heating directly, so it becomes very popular in areas where the
winter weather is mild. For AC and heat pump the “efficiency” can be much higher than 100%,
so a new terminology is used for them: coefficient of performance (COP).
Besides the applications in calculating the limit of efficiency of these systems, the second
law can also point out the direction of time in irreversible processes. For example, if we record
the process of a glass cup sliding off the edge of a table and shattering into pieces on the floor,
we could also play it backward: pieces of glasses are assembled into a glass cup on the floor and
then it rises to the top of a table. Everyone can tell which process cannot happen naturally in
the world: the rising up process violates the first law, and the self-assemble process violates the
second law due to reduction in entropy. Compared to the pieces of glass, a glass cup has much
lower entropy. Therefore, the arrow of time is pointing toward the direction of entropy increase
in isolated systems. Fortunately, we are not living in such a system: the earth receives energy
from the sun with the surface temperature at 6000 K, and it releases heat to the cold universe
background at the temperature of 2.725 K. If the earth is a heat engine, it can operate at a very
high efficiency.

2.4 THIRD LAW OF THERMODYNAMICS


The modern statement of the third law defines the reference point of entropy: “The entropy of a
perfect crystal at absolute zero degree is exactly equal to zero.” However, the classical statement
of this law is not so straightforward. The most popular one was the Nernst–Simon statement:
“It is impossible for any process, no matter how idealized, to reduce the entropy of a system to
its absolute-zero value in a finite number of operations.” This statement can be interpreted in
this way: as temperature approaches 0 K, the coefficient of performance of the heat pump also
approaches zero.
The operation of traditional heat pumps is based on two general characteristics of sub-
stances: the boiling point can be changed by varying the pressure, and the phase change process
requires a large amount of heat exchange. The details will be discussed in Chapter 3. However,
in many scientific experiments, temperatures below the boiling point of liquid helium (4.2 K) are
needed. An efficient way to achieve such a low temperature range is the adiabatic demagnifica-
tion process. In 2016, Cooltech Applications (a French company) was successful in developing
a refrigerator for commercial applications based on this process. Compared to traditional re-
frigerators, it can save up to 50% of energy and with much less noise and vibration. Figure 2.4
shows the principle of the cooling cycle.
20 2. THERMODYNAMICS LAWS
1 2
T T+∆T

4 3

T-∆T T

Figure 2.4: The cooling cycle of adiabatic demagnification refrigerator.

The mechanism in such a process is the magnetocaloric effect: when a susceptible material
is placed in a high magnetic field, it becomes magnetized, and the magnetic dipoles are aligned
along the direction of the external magnetic field. The work to rotate these magnetic dipoles
generates some heat that increases the temperature of the system. The heat generated can be
released to the environment easily, since its temperature gets higher. When the external mag-
netic field is removed or turned off suddenly, the reverse process happens: the magnetic dipoles
absorb heat from the system and their polarization directions are randomized, which lowers the
temperature of the system provided it is insulated from the environment.
This process can be analyzed more rigorously with the definition of entropy in microscopic
theory:
S D kB ln : (2.13)
In this equation, kB is the Boltzmann’s constant, and  is the number of states. Let’s assume
that the magnetic dipoles can be in two different orientations only, either up or down. When a
strong magnetic field is present, all of the magnetic dipoles are in the same orientation, so the
total number of states is  D 1, and the entropy of this magnetic subsystem is SM D 0. If the
magnetic field is turned off, each dipole can be in two states (up or down), and the total number
of states for N such dipoles are:  D 2N , and the entropy becomes: SM D N kB ln 2. Assume
this is a reversible process, and the amount of heat absorbed can be found:

Q D TSM D T  N  kB ln 2: (2.14)

We can see clearly that the amount of heat absorbed is proportional to the temperature. As the
system temperature approaches 0 K, the efficiency is also dropped to zero. In this way, we could
understand the Nernst–Simon statement.
For a perfect crystal at 0 K, the system is locked at the state with minimum energy, so
the number of state is  D 1, and the entropy is zero. However, if there is an impurity atom in
the crystal, the situation will be different. This impurity atom could be at the locations of any
2.4. THIRD LAW OF THERMODYNAMICS 21
atom in the crystal, and thus the number of states becomes 1 D N (the number of atoms).
Furthermore, with the rise of temperature, various vibration modes will be activated, which will
also generate many different states, so the entropy will rise rapidly with temperature.
In the case of ideal gas, the number of states can also be estimated. First, it depends on the
total number of particles:  D N 0 , where 0 is the number of states for each particle. In order
to simplify the analysis, we assume the particles are distinguishable. The entropy can be found
as S D N  kB ln 0 , which shows that the entropy is proportional to the number of particles.
This is consistent with the relationship between extensive and intensive properties, S D m  s ,
since the mass is proportional to the number of particles.
The number of states of each particle can be calculated in phase space, which is the com-
bined real space and momentum space. This can be illustrated with the example of a mass-spring
oscillator. Initially, the mass is pulled away from its equilibrium position and then released. As
a macroscopic object, its position and momentum can be measured at the same time. If friction
can be ignored, the measured data of location and momentum can be expressed as an ellipse,
which is shown in Fig. 2.5. At the beginning, it is located at the far-right end, where the dis-
placement is the greatest and the momentum is zero. When the mass comes to the equilibrium
position, it is at the bottom point of the ellipse: the displacement is zero, but with the high-
est magnitude of momentum. Therefore, the state of motion of the mass-spring oscillator can
be described by the elliptic trajectory moving in the clock wise direction. With friction taken
into account, the energy of the system will dissipate into heat gradually, and the trajectory in
the phase plane is expressed as a spiral moving inward. In this example, the phase space of the
mass-spring system is the area inside the ellipse.

o X

Figure 2.5: Phase space of a mass-spring oscillator.

In three-dimensional motions, imagine the particle as a small cube with finite volume V0 ,
and the number of locations in a container with volume VR can be found easily: NV D VR =V0 .
In momentum space, the situation is more complicated, and we can imagine that it is similar to
the situation of a planet with all its surface covered by ocean. When the temperature is low, the
surface of water is very calm, so the volume in momentum space is just the volume of the sphere.
However, when temperature rises, it is equivalent to the situation with huge waves rising above
the surface level and occupying the space above. Therefore, the volume in momentum space
22 2. THERMODYNAMICS LAWS
increases with temperature. Statistical physics provides an equation for calculating the number
of states in phase space:
VR VP
0 D ! S D kB ln  D N  kB ln 0 D N  kB .ln VR C ln VP 3 ln h/ : (2.15)
h3
In this equation, we find the number of states for each particle is proportional to both the volume
in real space .VR / and the volume in momentum space .VP /. Comparing it with the expression
of change in entropy in Eq. (2.10), we can see the consistency between microscopic and macro-
scopic definitions of entropy.
In fact, Eq. (2.15) is not accurate, and two other factors need to be considered. First,
the particles are indistinguishable. Second, uncertainty principle causes some correction. The
rigorous result is called Sackur–Tetrode equation:
   
VR 5
S D N  kB ln C : (2.16)
Nƒ3 2
p
In this equation a new parameter is introduced, the thermal length ƒ D h= 2mkB T , where m
stands for mass of a gas molecule.

2.5 MAXWELL RELATIONS


Some parameters in thermodynamics can be measured easily, while some others cannot. There-
fore, it is desirable to be able to find the relations between these parameters. Among the six
properties introduced, they can be classified into three groups: P and T in the first group are
pure intensive properties, and there are no extensive counterparts for them. The second group
includes v and s , which are complementary to P and T , respectively. As an analogy, if the first
group of properties play a role similar to force, then the second group of properties are equiva-
lent to displacement, and the product of them results in energy. This leads to the third group of
properties that include internal energy and enthalpy, which can be expressed as products of the
first two groups. For example, the First Law can be expressed in a differential format:
du D ıq ıw D T ds P dv ! u D u.s; v/: (2.17)
Similarly, the enthalpy, h D u C P v , can also be expressed in a similar way:
dh D T ds C vdP ! h D h.s; P /: (2.18)
Now we feel that something is missing: the combination of the four properties (T , P , v , s )
should result in four energy-related properties. Therefore, we need add two more properties:
Helmholtz function (a) and Gibbs function (g ). They were also called free energy in the past:
aDu Ts
(2.19)
gDh T s:
2.5. MAXWELL RELATIONS 23
Their differential relations can be derived easily:
da D sd T P dv ! a D a.T; v/ (2.20)
dg D sd T C vdP ! g D g.T; P /: (2.21)
The relationships can be summarized in Table 2.1.

Table 2.1: Differential relationships

P v
T g(T, P) a(T, v)
s h(s, P) u(s, v)

From these differential relations, the properties in the first and second group can be ex-
pressed as derivatives of the four energy-related properties (u; h; a; g ):
   
@u @a
P D D (2.22)
@v @v T
  s  
@u @h
T D D (2.23)
@s v @s P
   
@h @g
vD D (2.24)
@P s @P T
   
@a @g
sD D : (2.25)
@T v @T P
From these differential relations, we can also derive second-order relations. With mathematics,
we can derive the following relations in general:
   
@A @B @2 z
dz D A dx C B dy ! D D : (2.26)
@y x @x y @x@y

Therefore, the following relations can be derived:


   
@T @P
D
@v s @s
    v
@T @v
D
@P @s
  s  P (2.27)
@s @P
D
@v T @T v
   
@s @v
D :
@P T @T P
25

CHAPTER 3

Control Mass Processes


In many engineering applications, the working substance is confined in a closed container, so the
mass remains the same while other parameters are varying. In thermodynamics these are called
control mass processes. One example is the internal combustion engine, where the mixture of
air and fuel is confined in a cylinder, although the volume can change with the movement of the
piston.
From the beginning of this chapter, we will shift to non-ideal gases, where the relationship
between the properties cannot be expressed in simple equations. Therefore, we will rely on data
tables to find solutions. Although these tables look very messy, and the data were measured by
experiments, there are hidden relationships underneath. One could fit these data into analytic
equations, but it isn’t worthwhile to do so, since the equations would be very complicated.

3.1 STATE AND PROPERTIES


If you would like to describe the state of a satellite orbiting the earth, you need only to measure
two parameters: location and velocity. If the mass is known, you can derive many other related
parameters from these two observed parameters, such as potential energy, kinetic energy, lin-
ear and angular momentum, etc. On the other hand, if you know the kinetic energy or linear
momentum, you can also figure out the velocity. In a similar way, among the six properties (T ,
P , v , u, h, s ) that we discussed in the previous chapters, only two of them are independent.
Therefore, if you can find any two of them, the state is determined, and then you can find all the
other missing properties.
Figure 3.1 is the phase diagram of water in a T –P diagram. Suppose A is an arbitrary
point on this diagram; it can represent a unique state with two properties (TA , PA ). From these
two parameters, all the other properties (vA ; uA ; hA ; sA ) can be found by looking up data tables,
which are listed at the end of most textbooks on Engineering Thermodynamics. Due to limited
space in these tables, the data points are rather sparse. We can imagine that there is a grid in
this T –P diagram, where these data tables only list the properties at the intersections of the
grid lines. Fortunately, a number of software tools with computerized data tables are available,
such as CATT3 (Computer Aided Thermodynamics Tables 3). Ideally, any two properties can
be selected as input. Unfortunately, there are only eight options in CATT3, which is shown in
Fig. 3.2. If the temperature is known, one can select the options #1–4. Likewise, if the pressure
is known, options #5–8 can be used.
26 3. CONTROL MASS PROCESSES
P

Liquid
Solid

Gas

T
Ttr

Figure 3.1: Phase diagram of water in T –P diagram.

Figure 3.2: Input property sets in CATT3.

If neither T nor P is available, one can still find all the needed properties with an iterative
approach, as long as two of the six properties are known. For example, if one knows the specific
volume and enthalpy, then either option #2 or option #6 can be selected. Suppose option #2
with input of T and v is selected; initially, one needs to guess a temperature T1 , and the software
can find the corresponding enthalpy h1 . If it is lower than the known value of enthalpy, one can
choose a higher temperature T2 . This procedure might look a little tedious, but usually one can
reach the target in several trial and error iterations.
Figure 3.3 shows a sample data table for water in superheated vapor phase. The temper-
ature is listed in vertical direction in intervals of 50ı C. The pressure is arranged in a horizontal
direction; however, the limited width of the pages can only list the properties for two pressure
points. One can imagine that the original data table is displayed in a very long strip. In order
to show them in the standard pages in textbooks, this strip is cut into smaller segments with
only two pressure data points in each segment. Then they are rearranged by putting them one
3.2. PHASE CHANGE 27
Superheated Vapor Water
Temp. v u h s v u h s
(°C) (m3/kg) (kJ/kg) (kJ/kg) (kJ/kg-K) (m3/kg) (kJ/kg) (kJ/kg) (kJ/kg-K)

P = 10 kPa (45.81°C) P = 50 kPa (81.33°C)


Sat. 14.67355 2437.89 2584.63 8.1501 3.24034 2483.85 2645.87 7.5939
50 14.86920 2443.87 2592.56 8.1749 — — — —
100 17.19561 2515.50 2687.46 8.4479 3.41833 2511.61 2682.52 7.6947
150 19.51251 2587.86 2782.99 8.6881 3.88937 2585.61 2780.08 7.9400
200 21.82507 2661.27 2879.52 8.9037 4.35595 2659.85 2877.64 8.1579
250 24.13559 2735.95 2977.31 9.1002 4.82045 2734.97 2975.99 8.3555

Figure 3.3: Sample data table for water in the superheated vapor phase.

beneath the other. If one needs the data between these data points, a linear interpolation is often
needed.
Looking at these data tables, one should keep in mind that they are listed in families.
For example, the state with TA D 100ı C and PA D 50 kPa has four other family members:
vA D 3:41833 kg/m3 , uA D 2511:61 kJ/kg, hA D 2682:52 kJ/kg, sA D 7:6947 kJ/kg-K. Because
temperature and pressure are the two properties that can be measured easily, they are listed as
the primary parameters. However, one could create data tables with any two of the six properties
listed as primary parameters.

3.2 PHASE CHANGE


In many applications, the working substance undergoes phase changes in a cycle, such as in
power plants and refrigerators. In the T –P diagram shown in Fig. 3.1, the saturated state—the
boundary between liquid and gas—is represented in a line. It can be interpreted as a function
between pressure and temperature. In other words, there is a one-to-one relationship between
them. Therefore, if we know the temperature, we can find the corresponding pressure, and vice
versa. By comparing the saturated temperature or pressure, we can identify the phase with the
following criteria:
• Compressed liquid: P > Psat (above the line) or T < Tsat (to the left of the line).
• Saturated state: P D Psat and T D Tsat (on the line).

• Superheated vapor: P < Psat (below the line) or T > Tsat (to the right of the line).
However, the T –P diagram might be a little misleading. It seems that the phase transi-
tion can happen easily just crossing the boundary. Actually, a huge amount of heat exchange
is involved, so the T –v diagram in Fig. 3.4 can show this process in a better way. The liquid
phase is in the area to the left of the curve, and the vapor phase is in the area to the right side.
28 3. CONTROL MASS PROCESSES
Underneath the curve, it is in a saturated state with liquid and vapor coexisting at the same
time. Because the temperature remains constant in a phase change process, it is represented by
a horizontal line across the curve.
T

x=0 x=1

v
vf v vg

Figure 3.4: T –v diagram for a two-phase liquid-vapor region. (Courtesy of Wiley.)

For example, when water is heated to its boiling point, liquid water is gradually converted
into steam. If the steam is confined in a container with the liquid water while keeping the
pressure constant, we can observe that the component of steam rises and the component of
liquid water falls at the same time. Therefore, we need to define a parameter to describe the
percentage in mass of steam in such a mixture:
mg mg
xD D : (3.1)
mf C mg m

In this equation, mg stands for the mass of steam, and mf refers to the mass of liquid water.
This newly defined parameter x is called quality, which is meaningful only in the situation of
the two-phase mixture. Saturated liquid—meaning all the substance is in the liquid form—is
one extreme situation with x D 0. Similarly, saturated steam is the other extreme situation with
x D 1.
At atmospheric pressure, the boiling point of water is at 100ı C. Assume that the pres-
sure remains unchanged, and the water is heated from 50ı C to 150ı C. This process crosses the
boundary of liquid-gas phase transition. When the temperature is below 100ı C, it is in the
phase of compressed liquid, which is on the left of the curve in the T –v diagram, as shown in
Fig. 3.4. In this case, its specific volume is a little less than that of saturated liquid: v < vf .
When the temperature reaches 100ı C, it becomes saturated water (mixture of liquid water
and steam), and it will be stuck at this state for a long time. Initially, it is in the state of saturated
liquid (x D 0) and v D vf , which is shown at the left end of the horizontal line in Fig. 3.4. As
3.2. PHASE CHANGE 29
more heat is absorbed, some liquid water is converted into steam (0 < x < 1), and the specific
volume is between the values at the two ends: vf < v < vg . With more heat absorbed, it finally
reaches the state of saturated vapor (x D 1; v D vg ), which is shown at the right end of the
horizontal line in Fig. 3.4. After all the liquid water is converted into steam, the temperature
can rise again, and then it enters the phase of superheated vapor. In Fig. 3.4, it goes into the
region on the right side of the curve: v > vg .
Besides the T –v diagram shown in Fig. 3.4, one can replace the specific volume v with
u, h, or s , and these diagrams will have the similar configuration. Therefore, we can derive the
criteria for identifying the phase of a substance with given temperature or pressure, and the
parameter y stands for any of the four properties (v , u, h, s ):
• y < yf : compressed liquid,
• y D yf : saturated liquid,
• yf < y < yg : saturated water,
• y D yg : saturated vapor, and
• y > yg : superheated vapor.
Although the temperature and pressure remain the same in a phase change process, all the
other four properties change significantly. Figure 3.5 lists the specific volume and internal energy
for saturated water. We can see that the former one increases by three orders of magnitude in
the phase change process .vf ! vg /.
In many applications, we need to calculate the properties in the state of mixture with
0 < x < 1. For example, the specific volume can be found in this way:

V D Vf C Vg D mf vf C mg vg ! v D V =m D .1 x/vf C xvg :

The definition of quality was used in the derivation above. For convenience in calculation,
the table also lists the difference between vf and vg : vfg D vg vf . In this way, the specific
volume can be calculated in a very simple formula: v D vf C x  vfg . This is a general formula,
and it can also be applied in calculating internal energy, enthalpy, and entropy:

y D yf C x  yfg : (3.2)

As we know, an equation represents the relationship between variables. Therefore,


Eq. (3.2) can also be used in calculating the quality x if any of the four properties (v , u, h,
s ) is known.

Example 3.1
A 5-liter pressure cooker with 24 g of water is heated to 115ı C. Find its pressure and the
total internal energy of the water.
30 3. CONTROL MASS PROCESSES
Saturated Water
Specific Volume, m3/kg Internal Energy, kJ/kg
Temp. Press. Sat. Liquid Evap. Sat. Vapor Sat. Liquid Evap. Sat. Vapor
(°C) (kPa) vf vfg vg uf ufg ug
0.01 0.6113 0.001000 206.131 206.132 0 2375.33 2375.33
5 0.8721 0.001000 147.117 147.118 20.97 2361.27 2382.24
10 1.2276 0.001000 106.376 106.377 41.99 2347.16 2389.15
45 9.593 0.001010 15.2571 15.2581 188.41 2248.40 2436.81
50 12.350 0.001012 12.0308 12.0318 209.30 2234.17 2443.47
55 15.758 0.001015 9.56734 9.56835 230.19 2219.89 2450.08
90 70.14 0.001036 2.35953 2.36056 376.82 2117.70 2494.52
95 84.55 0.001040 1.98082 1.98186 397.86 2102.70 2500.56
100 101.3 0.001044 1.67185 1.67290 418.91 2087.58 2506.50
105 120.8 0.001047 1.41831 1.41936 440.00 2072.34 2512.34
110 143.3 0.001052 1.20909 1.21014 461.12 2056.96 2518.09
115 169.1 0.001056 1.03552 1.03658 482.28 2041.44 2523.72

Figure 3.5: Sample data table for water in superheated vapor phase.

The solution to this problem starts with identifying the phase of water in the pressure
cooker:
V 0:005
vD D D 0:208333 .m3 =kg/:
m 0:024
! Looking up the table in Fig. 3.5: vf D 0:001056 m3 /kg, vg D 1:03658 m3 /kg.
! vf < v < vg , it is in the phase of saturated water.
! P D Psat D 169:1 kPa.
v v
! x D vfgf D 0:208333 0:001056
1:03552
D 0:200.
! u D uf C x  ufg D 482:28 C 0:200  2041:44 D 890:57 (kJ/kg).
! U D mu D 0:024  890:57 D 21:374 (kJ).
Besides the internal energy, we can also find the enthalpy and entropy in the same way.

Example 3.2
At atmospheric pressure, 0.5-kg of water is heated from saturated liquid to saturated va-
por. Find how much heat is absorbed in this process.

There are two ways to solve this problem, because both pressure and temperature are con-
stant in this process. The data for enthalpy and entropy are shown in Fig. 3.6.
Solution A: With constant pressure, we can find the answer from change in enthalpy:

Q D m.hg hf / D mhfg D 0:5  2257:03 D 1128:52 (kJ):


3.3. ISOCHORIC PROCESS 31
Note: the parameter hfg is called latent heat for vaporization and condensation.
Solution B: With constant temperature, we can also find the answer from change in entropy:

Q D TS D mT .sg sf / D mT  sfg D 0:5  373:15  6:0480 D 1128:41 (kJ):

Saturated Water
Enthalpy, kJ/kg Entropy, kJ/kg-K
Temp. Press. Sat. Liquid Evap. Sat. Vapor Sat. Liquid Evap. Sat. Vapor
(°C) (kPa) hf hfg hg Sf Sfg Sg
0.01 0.6113 0.00 2501.35 2501.35 0 9.1562 9.1562
5 0.8721 20.98 2489.57 2510.54 0.0761 8.9496 9.0257
10 1.2276 41.99 2477.75 2519.74 0.1510 8.7498 8.9007
90 70.14 376.90 2283.19 2660.09 1.1924 6.2866 7.4790
95 84.55 397.94 2270.19 2668.13 1.2500 6.1659 7.4158
100 101.3 419.02 2257.03 2676.05 1.3068 6.0480 7.3548

Figure 3.6: Sample data table for water in superheated vapor phase.

3.3 ISOCHORIC PROCESS


As we introduced earlier, if a substance is confined in a rigid container with constant volume,
the process is called isochoric. In this situation, no work is done because there is no movement
of the boundary. From the First Law, the heat exchange is equal to the change of the internal
energy, since the work is equal to zero.

Q1!2 D U D m.u2 u1 /: (3.3)

Example 3.3
5 kg of water at 115ı C with a quality of 20% is contained in a rigid tank. When its
temperature is raised to 300ı C, find how much heat is absorbed in this process.

1. We can find the internal energy of the initial state, and the parameters at 115ı C can be
found in Fig. 3.5:

u1 D u1f C x1  u1fg D 482:28 C 0:2  2041:44 D 890:57 (kJ/kg):

2. We need to find the specific volume, which will remain the same in this process:

v2 D v1 D v1f C x1  v1fg D 0:001056 C 0:2  1:03552 D 0:20816 .m3 =kg/:


32 3. CONTROL MASS PROCESSES
3. Next we need to identify the phase of the final state by using the specific volume:

@ T2 D 300ı C; vf D 0:001404 m3 =kg; vg D 0:02167 m3 =kg:

Since v2 > vg , the final state is in the phase of superheated vapor.

4. Now we can find the internal energy of the final state with (T2 D 300 K and v2 D
0:20816 m3 /kg):
u2 D 2789 kJ/kg:

5. The amount of heat absorbed is equal to the change in internal energy (Eq. (3.3)):

Q1!2 D m.u2 u1 / D 5  .2789 890:57/ D 9492 (kJ):

3.4 ISOBARIC PROCESS


In many situations, the pressure of a system remains constant. Such an isobaric process can be
implemented in a cylinder-piston system. When heat is absorbed in an isobaric process, the
work done can be found very easily:

W1!2 D PV D mP .v2 v1 /: (3.4)

Likewise, the calculation of heat exchanged is also straightforward:

Q1!2 D H D m.h2 h1 /: (3.5)

The following problem is modified from Example 3.3, and the isochoric process is replaced
with an isobaric process.

Example 3.4
5 kg of water at 115ı C with a quality of 20% is contained in a cylinder-piston system,
where the presssure remains constant. When its temperature is raised to 300ı C, find how much
work it does and how much heat is absorbed in this process.

1. First, we can find the pressure at the initial state by looking up the table shown in Fig. 3.4.
There is a one-to-one relationship between pressure and temperature in the saturated state:

P2 D P1 D 169:1 kPa:

2. We can find the enthalpy of the initial state:

h1 D h1f C x1  h1fg D 482:46 C 0:2  2216:5 D 925:76 (kJ/kg):


3.5. ISOTHERMAL PROCESS 33
3. We need to find the specific volume of the initial state:

v1 D v1f C x1  v1fg D 0:001056 C 0:2  1:03552 D 0:20816 .m3 =kg/:

4. Now we can identify the phase of the final state:

T2 D 300ı C; Psat D 8581:0 kPa:

Since P < Psat , the final state is in the phase of superheated vapor.
5. We can find the enthalpy and specific volume of the final state with (T2 D 300 K, P2 D
169:1 kPa):
h2 D 3073 kJ/kg; v2 D 1:558 m3 =kg:

6. The work and the amount of heat absorbed can be found with Eqs. (3.4) and (3.5):

W1!2 D mP .v2 v1 / D 5  169:1  .1:558 0:20816/ D 1134:5 (kJ)


Q1!2 D m.h2 h1 / D 5  .30:73 925:76/ D 10736 (kJ):

7. We can verify the result with the First Law:

U D Q1!2 W1!2 D 9602 (kJ)


U D m.u2 u1 / D 5  .2809 890:57/ D 9592 (kJ):

Although these two results are not identical due to limited precision in the parameters,
they are very close to each other with percentage error of just 0.104%.
Comparing with the result of Example 3.3, we find that the isobaric process absorbs more
heat than the isochoric process, since a portion of the heat absorbed in the isobaric process is
converted to work.

3.5 ISOTHERMAL PROCESS


In a very slow process, the temperature in a system can be at equilibrium with the environment
and remains constant. In such an isothermal process, the heat transfer can be calculated from
the change in entropy, if it is reversible:

Q1!2 D TS D mT .s2 s1 /: (3.6)

With the First Law, the work can be found as well:

W1!2 D Q1!2 U D Q1!2 m.u2 u1 /: (3.7)


34 3. CONTROL MASS PROCESSES
Example 3.5
5 kg of water at 115ı C with a quality of 50% is contained in a cylinder-piston system,
then it expand slowly until its volume tripled. In this process the temperature remains constant.
Find how much heat is absorbed and the amount of work it does in this process.

1. We can find the internal energy, specific volume, and entropy of the initial state:
u1 D u1f C x1  u1fg D 482:28 C 0:5  2041:44 D 1503 .kJ/kg/
v1 D v1f C x1  v1fg D 0:001056 C 0:5  1:03552 D 0:51882 .m3 =kg/
s1 D s1f C x1  s1fg D 1:4733 C 0:5  5:7100 D 4:3233 .kJ/kg-K/:

2. We need to identify the phase of the final state:


v2 D 3v1 D 1:55645 m3 =kg
v2 > vg D 1:03658 m3 =kg ! the phase of the final state is superheated vapor.

3. Since the temperature (T2 ) and specific volume (v2 ) are known, all the other properties of
the final state can be found:
u2 D 2529 kJ/kg, and s2 D 7:38 kJ/kg-K:

4. The heat absorbed can be found from entropy change in isothermal process:
Q1!2 D mT .s2 s1 / D 5  .273 C 115/  .7:38 4:32/ D 5936:4 (kJ):

5. The work down can be found from the First Law:


W1!2 D Q1!2 m.u2 u1 / D 5936:4 5  .2529 1503/ D 806:4 (kJ):

3.6 ADIABATIC AND ISENTROPIC PROCESS


In a very fast process, the heat exchange with the environment can be ignored, and it can be
considered an adiabatic process. Following the First Law, the amount of work is equal to the
change in internal energy:
W1!2 D Q1!2 U D m.u2 u1 /: (3.8)
If it is a reversible process, the entropy is conserved, and it will become an isentropic
process: s2 D s1 .
Example 3.6
5 kg of water steam at 190ı C and 1 MPa is contained in a cylinder-piston system. Then it
expands rapidly until its volume quadrupled. Assume it is a reversible process. Find how much
work is done, as well as the temperature and pressure of the final state.
3.6. ADIABATIC AND ISENTROPIC PROCESS 35
1. First, we can identify the phase of the initial state:

T1 D 190ı C ! Psat D 1:2544 MPa: P1 < Psat ! superheated vapor.

2. We can find the internal energy, specific volume, and entropy of the initial state:

u1 D 2603 (kJ/kg)
v1 D 0:2003 .m3 =kg/
s1 D 6:642 (kJ/kg-K):

3. Next, we can find the specific volume and entropy in the final state:

v2 D 4v1 D 0:8012 .m3 =kg/


s2 D s1 D 6:642 (kJ/kg-K):

4. With these two state properties (v2 and s2 ), all the other properties of the final state can
be found with an iterative approach:

T2 D 120:5ı C; P2 D 202 kPa; u2 D 2354 kJ/kg:

5. The work done can be found from Eq. (3.8):

W1!2 D m.u2 u1 / D 5  .2354 2603/ D 1245 (kJ):


37

CHAPTER 4

Control Volume Processes


In many engineering applications, fluids are not confined in a container as in the control mass
systems discussed in the previous chapter. Instead, fluids flow through the devices with open
inlets and outlets; examples include heat exchangers, pumps, turbines, etc. For these open sys-
tems, our emphasis is on the imagined space inside the devices, and such a method is called
control volume (CV) analysis.

4.1 MASS RATE EQUATION


Most lakes have rivers flowing in and out of them. If the factors of precipitation, evaporation,
and seepage can be ignored, the water level of the lakes can be determined by the difference
between the inflow and outflow rates of the rivers. In many engineering devices the relationship
of mass can be analyzed in the same way:

d mC V X X
D m
Pi m
P o: (4.1)
dt
In this equation, mC V stands for the total mass of the fluid in the control volume, m
P i and m
P o are
the mass flow rates of the inflow and outflow, respectively. This equation reflects the conservation
of mass, which cannot be created or annihilated at this energy level.
In many engineering applications, the devices operate in a steady state, and the mass of
substance in the control volume remains constant. Therefore, we can have the following balance
equation:
X X
m
Pi D mP o: (4.2)

If there is a single inlet and a single outlet, such as with pumps and heat exchangers, this
equation can be further simplified:
mPi D mP o D m:
P (4.3)

4.2 ENERGY RATE EQUATION


As we know, besides mass conservation, energy is also conserved. However, this situation is a
little more complicated. When a fluid flows into a system, it does work by pushing itself into the
system. In the same way, when it flows out of a system, work is done to the fluid as it is pushed
out.
38 4. CONTROL VOLUME PROCESSES
Imagine the fluid is pushed in or pushed out by a piston, which is shown in Fig. 4.1, the
work done is equal to Wflow D F  L D P  A  L D P V D mP v , and the power can be found
as:
d
WP flow D .mP v/ D mP
P v: (4.4)
dt

Figure 4.1: Work at the inlet and outlet.

When fluids flow in or out of a device, they carry energy in or out of the system. The energy
carried with the fluids has two components—microscopic and macroscopic energy components:
   
1 2 1 2
E D mu C mv C mgz D m u C vm C gz : (4.5)
2 m 2
To distinguish velocity from specific volume, a subscript m is added to the symbol of the
velocity. In most situations, the velocity of fluid has a specific profile over the cross-section of a
pipe, and a typical profile of laminar flow in a circular pipe is shown in Fig. 4.2. The velocity vm
used in this equation is understood as the equivalent velocity for the kinetic energy of the real
flow.

Figure 4.2: Velocity profile in a circular pipe.

The total energy added to or subtracted from the system is the sum of the work and energy
carried by the fluid, and the total power can be found as:
    
P P P 1 2 1 2
Eflow D Wflow C E D m P Pv C u C v C gz Dm P h C vm C gz D mh P tot : (4.6)
2 m 2
2
In this equation, the total enthalpy is introduced as: htot D h C 0:5vm C gz . However, the units
of these three terms are different. In engineering thermodynamics, the unit of h is kJ/kg, but
4.3. NOZZLE AND DIFFUSER 39
the unit of the two terms of the macroscopic energy is J/kg. Therefore, for practical calculations
the following equation can be used:
2
vm gz
htot D h C C .kJ/kg/: (4.7)
2000 1000
For gases, the enthalpy h is usually much larger than the macroscopic energy, because the
speed of gas molecules is much higher than that of the macroscopic motion of the fluid. For
example, nitrogen molecules at room temperature has rms speed of 511 m/s, which is much
higher than the speed of sound at 343 m/s. Therefore, if the velocity of the fluid is not very
high (vm < 15 m/s) and the change in elevation is small (z < 10 m), one can safely ignore the
macroscopic energy terms. In that case, one can replace the total enthalpy with the conventional
enthalpy: htot  h.
Besides the energy associated with the inflow and outflow of fluids, the system can also
exchange heat with the environment, as well as deliver or receive power in the case of pump
or turbine. With all these factors taken into account, the following equation on energy can be
obtained:
dEC V X X
D QP C V WP C V C mP i htot;i mP o htot;o : (4.8)
dt
In a steady state, there is no change in the energy inside the control volume, so the left
side of the equation above vanishes:
X X
QP C V WP C V D mP o htot;o m
P i htot;i : (4.9)

In the situation with single inlet and single outlet, we can have the following simplified
equation:
QP C V WP C V D m.h
P tot;o htot;i / ! qC V wC V D htot;o htot;i : (4.10)
P m
In this equation, qC V D Q= P is called specific heat transfer, and wC V D WP C V =m
P is called specific
work.

4.3 NOZZLE AND DIFFUSER


The simplest CV devices are diffusers and nozzles, essentially, they are just pipes with changing
cross-section areas. However, they play important roles in many systems. For example, they are
at the front and the back ends of a jet engine, which is depicted in Fig. 4.3 in a simplified sketch.
The diffuser converts the high-speed air flow with low T and low P into a low speed air flow with
high T and high P . In other words, the macroscopic kinetic energy of the inflow is converted
into microscopic energy in the outflow. On the other hand, a nozzle operates in the opposite
way, and it converts the microscopic energy of the inflow fluid with high T and high P into
macroscopic energy of a high-speed jet stream with low T and low P .
40 4. CONTROL VOLUME PROCESSES

1 6
2 3 4 5
Diffuser Compressor Combustors Turbine Nozzle

Figure 4.3: Sketch of jet engine.

In a steady state operation, the mass flow rate is the same at the inlet and outlet: m Pi D
m
Po D mP . In addition, there is no moving component in the devices, so the work is equal to zero:
wC V D 0. Therefore, the following energy equation can be applied:
QP C V D m.h
P tot;o htot;i / ! qC V D htot;o htot;i : (4.11)

Example 4.1
Steam at 0.75 MPa and 215ı C enters an insulated horizontal nozzel with a velocity of
30 m/s. It leaves the nozzle at a pressure of 0.125 MPa and 110ı C, find the exit speed.

1. Because the nozzle is insulated, the heat transfer can be ignored.


htot;o D htot;i :

2. Since the nozzle is in horizontal position, the change in potential energy vanishes.
2 2
vm;o vm;i
ho C D hi C :
2000 2000
3. With pressure and temperature, the enthalpy can be found.
hi D 2876 kJ/kg; ho D 2694 kJ/kg:

4. The exit speed can be found with the parameters above.


2 2
vm;o D vm;i C 2000  .hi ho / D 3:649  105
vm;o D 604:1 (m/s):

To create such a supersonic jet, the nozzle needs to have a converging section and a di-
verging section, and it is called a de Laval nozzle, which is shown schematically in Fig. 4.4.
More details will be covered in the course of fluid mechanics.
4.4. COMPRESSOR AND TURBINE 41

Figure 4.4: A sketch of de Laval nozzle.

4.4 COMPRESSOR AND TURBINE


In a jet engine, both compressor and turbine are present, as seen in Fig. 4.3. Pump is another
name for compressor, which consumes power and increases the pressure of the fluid passing
through it. Turbines operate in the opposite way; they are engines that can export power derived
from the fluid passing through it. Besides jet engines, gas turbines are widely used as engines
for ships, tanks, helicopters, and turboprop aircrafts. In comparison with internal combustion
engines, jet engines or gas turbines have higher power-to-weight ratio, this is why they are
widely used in airplanes. In addition, water turbines are used in hydraulic power plants, and
steam turbines are used in conventional and nuclear power plants.
Besides these modern applications, turbines were also used in ancient times. For example,
a watermill is a form of primitive turbine, and it was used to process grains in villages for more
than 1,000 years. In addition, watermills also provided power for factories before the steam
engine was invented. Furthermore, windmills were invented 2,000 years ago in Greece, and
they were widely used in Netherlands since the 15th century A.D.
Some turbines have more than one inlet or outlet, so we need to use the general form of
energy conservation derived from Eq. (4.9):
X X
WP C V C mP o htot;o D QP C V C m
P i htot;i : (4.12)

In the special case of single inlet and single outlet, this equation can be simplified:
WP C V D QP C V C m.h
P tot;i htot;o /: (4.13)

Example 4.2
A steam turbine has one inlet and one outlet, and the mass flow rate is 1.25 kg/s. At the
inlet the steam is at 60 bar and 400ı C, and the velocity is 10 m/s. At the outlet, the pressure
drops to 1.1 bar and the quality is 92%, and the velocity is 30 m/s. The heat transfer rate to the
environment is 65 kW. Find how much power this turbine can deliver.

1. Find the enthalpy of the steam at the inlet with Pi D 60 bar and Ti D 400ı C:
hi D 3177 kJ/kg:
42 4. CONTROL VOLUME PROCESSES
2. Find the enthalpy of the steam at the outlet with Po D 1:1 bar and x D 0.92:

ho D 2500 kJ/kg:

3. Find the power:


" #
2
v 2m;i vm;o
WP C V D QP C V C m
P .hi ho / C
2000
 
102 302
D 65 C 1:25  .3177 2500/ C
2000
D 781 (kW):

4.5 THROTTLE VALVE


Throttle valves are used in many applications, and the one we use the most is a water faucet. In
addition, a throttle valve is also engaged while driving. When the accelerator of a car is pressed
down, the throttle valve attached to the intake manifold opens up to allow more air to come into
the engine. In general, when a fluid passes a throttle valve, there is a sudden drop in pressure.
As we know, the boiling point and pressure are positively correlated, and this is the basic
working principle of refrigerators. Figure 4.5 shows a simplified system diagram of a refrigerator,
which has four major components, and the refrigerant flows in the counter-clockwise direction.
The function of the compressor on the right is to increase the pressure, so the cool refrigerant
vapor condenses into warm liquid. The condenser is a heat exchanger outside the refrigerator,
and the refrigerant there is at a state with high pressure and high temperature. Therefore, heat
can be released into the ambient. The expansion valve (throttle valve) works in the opposite way
of the compressor, with the drop of pressure, warm liquid refrigerant evaporates into cool vapor.
The evaporator is the heat exchanger inside the refrigerator, and the refrigerant there is at a state
with low pressure and low temperature. In this way, heat can be absorbed from the foods stored
in the refrigerator. With the drive from the compressor, the flow of the refrigerant effectively
pumps the heat out of the refrigerator into the ambient.
In a throttle process, the heat transfer with the ambient can be ignored, and no work is
involved. Furthermore, the macroscopic kinetic energy terms can also be neglected. With all
these factors taken into account, the equation governing the throttle process is very simple:

ho D hi : (4.14)

Although this process is adiabatic, we cannot assume that the entropy is also conserved,
since it is an irreversible process.
4.5. THROTTLE VALVE 43

Q̇out
Condenser
3 2

Expansion
Valve Compressor Ẇin

4 Evaporator 1
Q̇in

Figure 4.5: Simplified system diagram of a refrigerator.

Example 4.3
Before the throttle valve, the refrigerant R-134a is in the phase of saturated liquid at 45ı C,
and the pressure drops to 200 kPa after going through the throttle valve. Find how much the
temperature drops in this process.

1. Find the enthalpy at the inlet with x D 0: hi D 264 kJ/kg, Pi D 1:16 MPa.

2. Find the enthalpy at the outlet: ho D hi D 264 kJ/kg.

3. Find the temperature of the final state with (Po ; ho ): To D 10ı C, x D 0:375.

4. Find the difference in temperature: T D Ti To D 55ı C.

Here we can ask a question: Does the temperature of gases always drop after a throt-
tle process? The answer is no, and it depends on a parameter, which is called Joule-Thomson
coefficient:  
@T v
J T D D .˛T 1/: (4.15)
@P h CP
 
In this equation, ˛ D v1 @T
@v
, which is the coefficient of cubic thermal expansion. If J T > 0,
P
the derivative of T over P is positive, so the temperature lowers after passing through a throttle
valve. Otherwise, the temperature rises. The curve with J T D 0 in Fig. 4.6 is the boundary
between the cooling region and the heating region in T –P diagram, and the scales are in re-
duced temperature and pressure: Tr D T =Tc and Pr D P =Pc , where Tc and Pc are the critical
temperature and pressure, respectively. Below this curve, J T > 0, and it is the cooling region.
44 4. CONTROL VOLUME PROCESSES

μJT < 0

Reduced Pressure (Pr)


μJT > 0

Reduced Temperature (Tr)

Figure 4.6: Regions with different signs of Joule-Thomson coefficient.

4.6 ENTROPY RATE EQUATION


Following the same approach of mass and energy analysis, we can find a similar relationship of
entropy. However, unlike matter and energy, entropy can be generated in irreversible processes.
Fortunately, in many situations, the entropy generation rate is not very high and thus can be
ignored. In addition, besides the entropy carried by the incoming and outgoing fluids, the heat
transfer with the environment also causes entropy change in the system:
dSC V X X QP C V
D m
P i si m
P o so C : (4.16)
dt T
In a steady state, the time derivative on the left side vanishes. For the situation with only
one inlet and one outlet, this equation can be simplified:
QP C V D mT
P .so si / ! qC V D T .so si /: (4.17)
In this situation, we can derive the power formula of a CV system. First, imagine that
the flow from inlet to outlet can be divided into many small steps, and then we can find the
differential forms of Eqs. (4.10) and (4.17):
ıwC V ıqC V D dhtot
(4.18)
ıqC V D T ds:
Combining these two equations, we get the following formula:
ıwC V D T ds dhtot : (4.19)
4.6. ENTROPY RATE EQUATION 45
2
Since htot D h C 0:5vm C gz and dh D T ds C vdP (Eq. (2.17)), the differential work is:
2

ıwC V D vdP d 0:5vm C gz : (4.20)
With an integral, the specific work exported from a CV system can be found:
Z o
1 2 2

wC V D vdP vm;o vm;i g .zo zi / : (4.21)
i 2
Multiplied by the mass flow rate, the exported power can be obtained: WP C V D mw
P CV .
Equation (4.21) is different from the formula of a CM system (e.g., cylinder-piston sys-
R2
tems), wCM D 1 P dv . The first term of the specific work in Eq. (4.21) is the negative integral
over pressure, which is shown graphically in Fig. 4.7. If the inlet pressure is higher than the out-
let pressure, the power exported is positive, such as in the situation of the turbine. However, in a
pump or compressor, the situation is opposite, so the specific work is negative, which indicates
that external power is needed to drive the device.

Figure 4.7: Work from a CV system.

Example 4.4
The mass flow rate of a water pump is 0.25 kg/s. At the inlet, the water is at 101 kPa
and 30ı C. At the outlet the pressure increases to 3 MPa. Assume the process is isentropic, and
the changes in velocity and elevation can be ignored. Find how much power is needed for this
pump.

1. Find the specific volume of water with Ti D 30ı C and Pi D 101 kPa:
v D 0:001004 m3 =kg:

2. Because the specific volume of water can be assumed constant, the specific work can be
found easily:
Z o
wC V D vdP D v.Pi Po / D 0:001004  .101 3000/ D 2:91 (kJ/kg):
i
46 4. CONTROL VOLUME PROCESSES
3. Find the power:

WP C V D mw
P CV D 0:25  2:91 D 0:728 (kW):

As we see in this example, for incompressible fluid (v or  is a constant), the integral


in Eq. (4.21) can be simplified. If it is applied to a free flow situation without power input or
output, we can derive the Bernoulli equation:
   
1 2 1 2 1 2 P
vm C gz C P v D vm C gz C P v ! vm C gz C D const: (4.22)
2 o 2 i 2 

To be consistent, the unit of pressure in this equation is Pa, instead of kPa.


Figure 4.8 shows an incompressible fluid flowing through a pipe with changing cross-
section area. Applying Bernoulli’s principle in this situation, where the elevation z is the same:
   
1 2 1 2
vm C P v D vm C P v : (4.23)
2 1 2 2

Due to mass conservation, we have the following continuity relationship of the volume
flow rate:
A1 vm;1 D A2 vm;2 : (4.24)
Therefore, the velocity at the section of the pipe with a smaller cross-section area is higher,
and thus the pressure there is lower.

Figure 4.8: Application of Bernoulli’s principle.

Rigorously speaking, the velocities in Eqs. (4.23) and (4.24) are not the same. In the
Bernoulli’s equation the rms velocity is used, which is related to the kinetic energy. On the other
hand, the average velocity is used in the continuity equation. Therefore, it is an approximation in
the analysis above, although the conclusion is correct. In general, Bernoulli’s principle is applied
along a streamline of the fluid.
47

CHAPTER 5

Advanced Topics
The theory of thermodynamics is not limited to physics and engineering. Instead, it can be ap-
plied in many areas. In this chapter, we will discuss a few interesting topics, such as information
theory, chemical reactions, dissipative structures, economy, etc.

5.1 ENTROPY AND INFORMATION


Information stored in electronic devices is in the binary form; in other words, there are only
two symbols: 0 and 1. For example, the core of a DRAM (Dynamic Random Access Memory)
cell is a capacitor, which could be charged or discharged. A capacitor is a container of electric
charges, so there are two distinct states: full or empty. The corresponding voltages are high (H )
and low (L), which can be represented by binary numbers 1 and 0, respectively. Actually, such a
capacitor is a leaky container needing frequent refreshing because electrons stored there will leak
out rather quickly. Each DRAM cell can store 1-bit information, so the amount of information
can be calculated in this way:
H D log2 N (bit): (5.1)

In this equation, N stands of the number of possible states. More rigorously speaking, it is better
to call it the amount of data instead of information.
If there are two DRAM cells, there are four outcomes: HH (11), HL (10), LH (01),
LL (00). In this case, each outcome carries 2 bits of information, which is derived from H D
log2 4 D 2. Furthermore, if the number of cells is n, then the number of outcomes is N D 2n .
In this case, each outcome can carry n-bit information, since H D log2 2n D n.
Another example is the genetic information stored in DNA, where four different base
pairs are available. According to this formula, each base pair can carry 2 bits of information,
provided that the probability or frequency of occurrence is the same. Human genome contains
roughly 3 billion base pairs, so the amount of information is about 6 billion bits or 750 MB
(1 Byte D 8 bit), which is close to the capacity of a regular CD with the capacity of 700 MB.
According to Eq. (5.1), the more the number of states, the larger the amount of informa-
tion that can be represented. However, detecting the information is another story. For example,
we could use one capacitor to store 3 bits of information by setting up 8 different voltage levels.
It seems like a good idea, but it is very hard to implement reliably. Therefore, besides the stor-
age capacity of information, one should also consider the reliability of the information retrieval
process.
48 5. ADVANCED TOPICS
From the definition of entropy, S D kB ln W , we can see the similarity with the definition
of information. In the same way, entropy is also closely related to the number of states of the
particles. If some particles stick together, the number of states of them will reduce significantly.
This principle can be applied in information compression. For example, in a black and white
digital picture the pixels are just like the particles. If there are patterns in the image, such as a
black polygon in the white background, the pixels at the boundary can be used to describe the
shape of this object. Actually, it can be further reduced to the coordinates of the vertices, since
an edge can be determined by the two corner points. In this way, the amount of information
describing such an image can be reduced substantially.
In many situations, we are provided with partial information, and then the uncertainty
and the amount of information is reduced. For example, the weather forecast tells us that the
probability of rain tomorrow is 10%, which provides some information on the weather. The next
morning, we find the outcome of the weather condition (sunny or rainy). How much information
does it contain in general?
We need to rewrite Eq. (5.1) in a way that can be related to the probability. In an unbiased
situation, the probability of each outcome is equal to the inverse of the number of possible
outcomes: p D 1=N or N D 1=p . Therefore, for each outcome, the amount of information can
be calculated in the same way.
1
Raining: pR D 0:1; HR D log2 pR
D 3:32 (bit) (unexpected result, high in information).
1
Sunny: pS D 0:9; HS D log2 pS
D 0:152 (bit) (expected result, low in information).
Average Information: H D pR HR C pS HS D 0:469 (bit).
Without the partial information provided from the weather forecast, pR D pS D 0:5,
the amount of information of the outcome becomes 1 bit. Therefore, the partial information
provided by the weather forecast can reduce the amount of information in the outcome.
The relationship between probability and information is shown in Fig. 5.1. The horizontal
axis indicates the probability of one of the two outcomes, which is the pre-knowledge from other
sources, and the vertical axis represents the amount of the information of the outcome. From
a certain point of view, the amount of information is related to our ignorance. In the case of
complete ignorance, the amount of information is the highest at 1 bit.
We can generalize this formula by calculating the amount of information in situations
with any number of discrete outcomes:
N
X N
X
1
H D pi log2 D pi log2 pi : (5.2)
pi
i D1 i D1

For example, the English alphabet has 26 letters. If the probability of appearance is the
same, each letter can carry 4.7 bits of information. However, the frequency of appearance of
the letters is very different, so the amount of information carried in each letter is lower. This is
5.1. ENTROPY AND INFORMATION 49
H
1.0

0.5

0 PR
0 0.5 1.0

Figure 5.1: Information vs. probability.

similar to the situations of unstrained expansion of gas and diffusion in liquid, where the uniform
distribution case has the maximum entropy.
When Claude Shannon was investigating the capacity of communication channels in
1948, he postulated Eq. (5.2) as the formula to calculate the amount of information. A sim-
ilar formula was proposed ten years earlier by John Slater in calculating entropy:
X
S D kB fi ln fi : (5.3)
i

In this equation, fi stands for the probability of the state i , which is equivalent to the pi in
Eq. (5.2). By comparing these two equations, we can see the connection between entropy and
information.
Manipulating information requires consumption of energy, no matter how it is imple-
mented. A primitive information storage device was a rope, where information was recorded by
tying knots. From a certain point of view, information can be interpreted in a different way:
Information D In C Formation. Now we can store information by creating formations in many
ways, such as in electronic, photonic, and magnetic devices. Creating a new formation of mat-
ter requires energy; in 1961, Landauer found the lower limit of energy to manipulate 1 bit of
information:
E1 (bit) D TS D kB T ln 2  0:693 kB T: (5.4)
Because the Boltzmann constant is very small .kB D 8:62  10 5 eV/K/, this lower limit
of required energy is also very small. For example, at room temperature (T D 300 K), this min-
imum required energy is equal to 0.179 eV. Therefore, information can be manipulated very
efficiently.
Intelligence—the capability of processing information—allows animals and humans to
obtain resources more effectively, otherwise the brain would become a burden. For example, the
human brain takes about 2% of body mass, but it consumes about 20% of energy. In many en-
gineering applications, information can also help us to obtain energy and power. An interesting
50 5. ADVANCED TOPICS
example is the tidal power generator, and its mechanism is very simple: at high tides, water is
stored in a reservoir, and it is released during low tides. The flow of water can turn a hydraulic
turbine and generate power. Similarly, the traders in stock market make money also in this way.
In 1966, the first large-scale tidal energy harnessing power plant, Rance Tidal Power Station,
started to operate in France, which generated 240 MW power at a lower cost than nuclear power
plants.
Although the second law of thermodynamics shattered the wonderful dream of perpetual
machines, many people still tried to design them with various clever ways. One well-known
example engaged “Maxwell’s demon.” As shown in Fig. 5.2, a container is separated into cham-
bers A and B by a membrane, which has a magic valve controlled by a demon (not shown). It
allows molecules with high kinetic energies to pass through the membrane from A to B. At the
same time, it allows molecules with low kinetic energies to pass through the membrane from
B to A. After a while, chamber A will have a lower temperature than chamber B. In this way,
we can obtain work or power for free by connecting a heat engine to these two chambers, and a
perpetual machine is constructed. Of course, such a design is flawed, since the demon also needs
energy input for doing the job.

A B

Figure 5.2: Two chambers separated by a magic valve.

5.2 EQUILIBRIUM IN ISOLATED SYSTEMS


The simplest situation is an isolated system, in which there is no exchange of matter, energy, or
information with the environment. In this case, the criterion for equilibrium is that the entropy
has achieved its maximum value:
S ! Smax : (5.5)
With this criterion, we can tell how an isolated system evolves, since the equilibrium state is
the destination. Figuratively speaking, such a process shows the arrow of time. For example, as
we discussed in Section 2.2, if two objects with different temperatures are put into contact, heat
will flow from the object with the higher temperature to the one with the lower temperature.
This process will keep going until the temperature difference disappears, when the entropy of the
5.2. EQUILIBRIUM IN ISOLATED SYSTEMS 51
system has achieved the maximum value. In the example of unstrained expansion, gas molecules
will eventually occupy all the space in an isolated container, because the entropy becomes largest
in this situation. In addition, if a drop of orange juice is dropped into a cup of water, it will diffuse
spontaneously and becomes uniformly distributed. From these examples, we can summarize that
the homogeneous state has the highest entropy.
Mathematics shows that the uniform distribution of particle is not always the state with
maximum entropy, and it is valid under two conditions. First, the potential energy is the same
everywhere. Second, there is no interaction between the particles. The examples of unstrained
expansion and diffusion in liquid satisfy both conditions, hence the uniform distribution is the
equilibrium state. However, our atmosphere is not in such a situation, since the gravity pulls
the air molecules down toward the earth. Near the surface of the earth, the gravity force can be
assumed constant. In this case, the density of air decreases as an exponential function with the
height. In addition, the water molecules in the atmosphere are not uniformly distributed either,
and they tend to stick together to form cloud because of the hydrogen bonds between these
molecules. In the universe, the stars were formed from dust clouds, where matter was gathered
by the gravitational force.
Even in the perfect situation without potential energy and interaction, from time to time
the distribution of particles can also deviate from the idealized homogeneous distribution. In
Fig. 5.3, a container is virtually divided into two identical chambers; in this way we can find the
probability with all particles located in one chamber. Assuming the particles are distinguishable,
the calculation is straightforward. For each particle, the probability for it to be in either chamber
is 0.5. Therefore, the probability of all the N particles located in one chamber is pN D 2 N . If the
particle number is rather small, this probability is not negligible. For example, with N D 4; p4 D
6:25%. For comparison, the probability of equal distribution is p2 2 D 37:5%. However, if the
number of particle is in the order of magnitude of Avogadro’s number, the probability of the
extreme case is completely negligible.

Figure 5.3: Particle distribution in a two-chamber system.

Although it is unlikely for all the particles to be crowded into one chamber, it is not
necessary to have exactly the same number of particles in the two chambers all the time. At a
finite temperature, the particles are moving about constantly, and the distribution of particles
is also varying, and small deviations happen all the time: NL D 0:5N N and NR D 0:5N C
N . As jN j increases, the probability decreases rapidly.
52 5. ADVANCED TOPICS
Figure 5.4 shows the probability distribution of ten particles in such a two-chamber sys-
tem. First, it shows that the situation with equal partition of particles .5  5/ has the highest
probability. Actually, there is no mystery about this; instead, it is just a result from mathemat-
ics: p.N; k/ D C.N; k/=2N , where C.N; k/ stands for the combination value. With equal par-
tition, there are more combinations of the particles than the other partition schemes. Second,
when k D N=2 ˙ 1, the probability is just slightly less than that in the equal partition situa-
tion, and the difference decreases when the particle number gets larger. Third, the probability
distribution is symmetric, so the average numbers of particles in both chambers are the same:
< NL >D< NR >D N=2. In other words, if measured for a long time, the average of the offset
number is zero: < N >D 0.
0.25
0.2
0.15
0.1
0.05
0
0~10 1~9 2~8 3~7 4~6 5~5 6~4 7~3 8~2 9~1 10~0

Figure 5.4: Probability distribution of ten particles in a two-chamber system.

An electric conductor can be viewed as a container of free-moving electrons, and it is


just like the situation of gas molecules confined in a container. Because electrons carry electric
charges, their motion will give rise to electric current. If a circuit is constructed with conductors,
a random current can be measured. This will cause interference to sensitive instruments, and
it is called thermal noise. If the noise power is measured, it turns out to be proportional to the
temperature in Kelvin. Since the probability distribution is symmetric, thermal noise can be
significantly reduced by averaging, if the signal is periodic. Most digital oscilloscopes have this
function, which can clean up weak and noisy signals effectively.

5.3 EQUILIBRIUM WITH HEAT TRANSFER


Isolated systems are very rare in the world, and the exchange of energy with the environment
by means of heat transfer is very hard to prevent. In such situations, change can happen in
both ways. For example, when one is served with a cup of water with ice in a restaurant, the
temperature of the water is at 0ı C. If it is left on a table at room temperature, the ice in the water
will melt gradually. However, if the cup is placed in the freezer of a refrigerator, it will freeze
into a chunk of ice. In the same way, many chemical reactions are reversible—depending on the
ambient conditions. Therefore, we need to investigate the relationship between the equilibrium
state and the factors of the external environment.
5.3. EQUILIBRIUM WITH HEAT TRANSFER 53
Figure 5.5 shows that we could include the system and its surroundings into a larger sys-
tem, which can be considered an isolated system. Now the original system becomes subsystem A,
and its surroundings becomes subsystem B .

Subsystem A

Subsystem B

Figure 5.5: A system and its surroundings form an isolated system.

Consider a chemical reaction in the subsystem A:

X C Y ! X Y: (5.6)

There are two different situations in this synthesis process: it could absorb heat (endother-
mic) or release heat (exothermic). As we discussed in Section 2.4, the entropy will decrease when
the particle number is reduced in situations like this synthesis reaction. If this reaction in the
subsystem A absorbs heat, the surroundings (the subsystem B ) will lose heat, and its entropy will
also decrease. Therefore, the entropy of the whole system will decrease, and then it is impossible
for it to proceed spontaneously. This is similar to the process of freezing or condensation, and
it is impossible for this to happen if it absorbs heat instead of releasing heat.
How about a heat-releasing synthesis reaction? Does it always happen spontaneously? In
this case, the entropy of these two subsystems changes in opposite directions. The reduction in
particle number reduces entropy in the subsystem A. On the other hand, the heat released in
the chemical reaction is absorbed by subsystem B , and its entropy is increased. Therefore, we
need a parameter that can take into account both of these factors. If it is in a rigid container
with constant volume, this parameter is the Helmholtz function or free energy: A  U T S .
Likewise, if the pressure remains the same, it is the Gibbs function or free enthalpy: G  H
T S . The criterion of equilibrium in such situations is the minimization of these parameters:

A ! Amin ; G ! Gmin : (5.7)

Most chemical reactions take place in an environment with constant pressure, so we will
concentrate on Gibbs function in this situation. The heat exchange is associated with enthalpy
54 5. ADVANCED TOPICS
change: ıQ D dH . Therefore, the differential change of Gibbs function can be expressed as:
dGA D dHA T dSA D ıQA T dSA : (5.8)
Since ıQA D ıQB D T dSB , the change in Gibbs function in subsystem A can be related to
the entropy change of the whole system:
dGA D T .dSB C dSA / D T dStotal : (5.9)
Therefore, the minimization of the Gibbs function in subsystem A is equivalent to the maxi-
mization of the entropy of the whole system. A similar equation can be derived with Helmholtz
function for a subsystem in a rigid container with constant volume.
Unlike the isolated systems, temperature plays an important role in this case, which can
be found in Eq. (5.8). This is also illustrated in Fig. 5.6: the temperature—equivalent to the
fulcrum position—determines which term has the upper hand. When the temperature is low, the
second term .T dSA / is low, so the effect of entropy change in the subsystem A is less important.
Therefore, the heat-releasing process decreases the Gibbs function, so the synthesis reaction
could happen. On the other hand, when the temperature is high, the decrease of entropy in
subsystem A could dominate the term of heat released. In this way, the Gibbs function will
increase, and the reaction will not happen. In summary, the synthesis reaction is likely to happen
at low temperature, but not at high temperature. This is similar to the phase change processes
of water: lower temperature favors freezing, and higher temperature causes evaporation.

Entropy Energy

Temp

Figure 5.6: Competition between energy and entropy.

Liquid water is also an interesting example, since the H2 O molecules can form “struc-
tures” with hydrogen bond, which is a weak interaction between the water molecules. Many
strange behaviors of water can be understood from this point of view. For example, the coef-
ficient of thermal expansion becomes negative when the temperature is between 0–4ı C. As a
simplified model, one can imagine that there are tiny ice structures in liquid water, and these
structures grow gradually when it cools down in this temperature range. Because the structured
configuration occupies larger space than the unassociated molecules, the volume will increase as
the temperature lowers. This can be understood easily with the analogy of Legos, since the same
set of pieces with structures formed cannot be put back into the original packaging box.
5.3. EQUILIBRIUM WITH HEAT TRANSFER 55
All the discussions above assumed that the Gibbs function changes monotonically during
the chemical reaction, so it either takes place spontaneously or is impossible to happen, depend-
ing on whether heat is absorbed or released. However, most situations are not that simple. Fig-
ure 5.7 shows the cases with a peak (barrier) in Gibbs function between reactants and products.
Therefore, even heat-releasing exothermic reactions do not necessarily take place spontaneously,
since they need some energy to overcome the barrier. Many combustion processes give off heat,
but such chemical reactions will not happen spontaneously. For example, one can ignite a can-
dle with the flame of a match. The ignition process provides the energy required to overcome
the barrier, and the heat released in the combustion process will keep the reaction going on. In
addition, one can extinguish a candle by blowing the flame off, and the chemical reaction is shut
down by removing the activating heat. In the same way, wind blower is the tool of choice in
fighting wild fires, since it is unrealistic to extend water pipe to remote areas.

G G

P R

∆G ∆G
R P

(a) Time (b) Time

Figure 5.7: (a) Endothermic reaction and (b) exothermic reaction.

Although an endothermic synthesis reaction is unfavorable from the criterion of Gibbs


function, it could occur when sufficient external energy with low entropy is provided. Photo-
synthesis is such an example, where CO2 and H2 O are synthesized into carbohydrate or sugar
while releasing oxygen: 6CO2 C 6H2 O ! C6 H12 O6 C 6O2 . In this case, the external energy
source is the sunlight, which provides the needed energy to overcome the barrier as well as lands
the product on a higher ground on Gibbs function.
Nuclear fusion reactor has been long proposed as the future of energy source, but it has
encountered tremendous technical challenges. If two deuterium atoms—an isotope of hydrogen
with one proton and one neutron—can be fused into a helium atom, it will release a lot of en-
ergy, which is just like an exothermic synthesis reaction. However, since there is a very strong
Coulomb force repelling the positively charged deuterium nuclei away from each other, an ex-
tremely high barrier of free energy is formed in between. Figure 5.8 shows the potential energy
between two charged nuclei, where the action of strong attractive nuclear force is effective only
56 5. ADVANCED TOPICS
in a very short distance ( 10 15 m). Inside the sun, this barrier is overcome by the tremendous
high pressure caused by the gravity.

PEtot

Repulsive
Coulomb
0 r

Attractive Nuclear

Figure 5.8: Potential energy over the distance between charged nuclei.

On the other hand, fission reactors are much easier to construct, since the entropy increases
when the nuclei split up. In addition, there are some very unstable isotopes, such as U235 , and
the fission process can be triggered by the impact of a neutron. However, the danger of fission
reactors is a runaway reaction that causes the core to melt down. Such disasters happened in
Chernobyl nuclear power plant in 1986, as well as in Fukushima Daiichi power plant in 2011.
A fission nuclear reaction is similar to a chemical decomposition process:

X Y ! X C Y: (5.10)

Since the number of particle increases, the change of entropy in subsystem A is positive.
If heat is released in the process, the entropy of the surrounding will also increase, so it could
happen spontaneously.

5.4 CRYSTAL GROWTH


From the point of view of chemistry, everything is just a combination of atoms in different
configurations. When the temperature gets higher, the bonds between atoms will get loose,
and they can disintegrate at a sufficiently high temperature. At even higher temperatures, the
electrons can also be freed from the atoms, and a state of plasma is formed. By manipulating
the temperature, we can dissolve old structures and form new ones.
This principle can be applied in casting process in metallurgy. For example, when a car
comes to the end of its life, it will be compressed into a block and then put into a smelting
furnace. The high temperature there transforms metals from solid phase into liquid phase, which
5.4. CRYSTAL GROWTH 57
will be poured into molds to form new structures when they are cooled down. From the point
of view of thermodynamics, the heat absorbed in the furnace increases the entropy and dissolves
the original structure. On the other hand, the decrease of entropy in the cooling process enables
it to keep the new form.
The same principle can be applied in crystal growth processes. Figure 5.9 shows a diagram
of molecular beam epitaxy (MBE) system, which is widely used in research labs to grow various
semiconductor devices in the way of layer by layer at atomic precision. Although the system
is rather complicated, the basic principle is very simple: atoms are emitted from the sources
at a high temperature and strike the substrate at a lower temperature, and then bond into the
substrate and form crystal structures.

Vacuum Pump

Cryo-panel

Mass Spectrometer

Substrate
RHEED Gun

Effusion Cells

Figure 5.9: Molecular beam epitaxy system.

There are three components at different temperatures in an MBE system. The first one is
the source of the materials, which is called effusion cells. Traditionally, solid sources are used,
and the atoms can be freed when heated at very high temperatures. For example, growing
GaAs/AlGaAs heterojunction or superlattice needs three effusion cells, which hold Ga, Al,
and As sources in them separately.
The second component is the substrate where the atoms are deposited, and it provides the
template of the crystal structure. In order to grow the crystal uniformly, some mobility is needed
for the atoms arriving at the surface of the substrate, so the sample holder is heated to a certain
temperature. This temperature needs to be controlled very carefully. If it is too low, the mobility
of arrived atoms is very low, and tiny islands will be formed at the surface. On the other hand,
if the temperature is too high, the arrived atoms cannot form bonds with the substrate, just like
the synthetic chemical reactions we discussed in the previous section.
58 5. ADVANCED TOPICS
The third component is the inner surface of the chamber, which is cooled by liquid ni-
trogen. In this way, the strayed atoms arrived there will be stuck and locked up. Usually, MBE
systems are used to grow different kinds of semiconductors, so various atoms are present on the
inner surface. If the temperature there is not cold enough, the incoming atom can strike out the
deposited atoms, which could land on the growing sample causing severe contamination.
In this example, we can see that temperature is related to freedom. The art of growing
good crystals with MBE system is controlling the freedom of the atoms. The idea of growth
on a template can be extended to biological systems. It is believed that the building blocks of
the most primitive life forms were RNA molecules, which can be self-replicated. Gradually,
RNA evolved in two directions: DNA provided the genetic code that can be replicated more
reliably, and protein molecules became the building blocks that have many more functions. In
a biological system, the information carried in DNA is transferred to RNA first, and later it is
used as a template to configure different kinds of protein molecules. In other words, the template
provides a low energy state for a specific protein molecule, so the atoms can be fused together
with decreasing entropy. Here, the trade-off between energy and entropy is demonstrated again.

5.5 CHEMICAL POTENTIAL


Many equations discussed in the first four chapters assume the number of particles remaining
the same. When this restriction is relaxed, these equations need to be modified. For example,
during phase change, molecules can leave one phase and join another phase, so the number of
molecules in a certain phase is no longer constant. In diffusion processes, solute molecules can
move in or out of a region, so the number of the molecules in a certain region also changes. In
general, when particles move into or out of a system, they carry energy with them, as well as
entropy. Let’s investigate the internal energy in this situation:

d U D T dS pd V C dN: (5.11)

In the new term added, dN stands for the change of particle number, and  is defined as
chemical potential (CP) with the dimension of energy. In this equation, (T , p , ) are the three
parameters that determine the spontaneous flow or movement directions of heat, boundaries,
and particles. Just like heat will transfer spontaneously from a high temperature region to a low
temperature region, particle will move spontaneously from a higher CP region to a lower CP
region. In the example of diffusion, the region with high concentration of a solute has a higher
CP, so the direction of diffusion is toward the regions with a lower concentration of the solute.
From Eq. (5.11) the chemical potential can be defined:
 
@U
D : (5.12)
@N S;V
5.5. CHEMICAL POTENTIAL 59
Keeping entropy (S ) and volume (V ) constant is not very convenient, so it is more popular
to engage Gibbs function:
dG D Sd T C Vdp C dN: (5.13)
In this way, the chemical potential can be understood as the increase of Gibbs function
by adding one particle into the system while keeping the temperature and pressure constant:
 
@G
D : (5.14)
@N T;P
If we allow one of the two parameters (T and P ) to change, something interesting can
happen. First, let the pressure remain constant and change the temperature, and then the second
term in Eq. (5.13) can be removed:
dG D Sd T C dN: (5.15)
From this equation we can find the slope of chemical potential versus temperature, which
has the same format of Maxwell’s equations:
   
@ @S
D < 0: (5.16)
@T p;N @N p;T
@S
Since entropy will increase with particle number, @N > 0, the derivative of chemical po-
tential over temperature is negative. In other words, the function of chemical potential vs. tem-
perature has a negative slope. In addition, the gradient of the slope is quite different in various
phases, which is related to the freedom of the molecules. For example, in the solid phase the
molecules are bonded together, so the entropy change is very little by adding one molecule to
it. On the other hand, molecules can move freely in the vapor phase, and then the change of
entropy with an additional molecule is much higher than in the solid phase.
Figure 5.10a shows a sketch of chemical potentials of the three different phases, and the
intersection points indicate the phase transition. Since a lower chemical potential is preferred,
this diagram explains the direction of phase change at different temperatures, which is equivalent
to the criterion with Gibbs function. In winter, salt is sprinkled on the roads, which reduces the
freezing point of water. In general, when solute is added to water, the chemical potential line
of the liquid section becomes lower, which is indicated as the dashed line in Fig. 5.10a. On the
other hand, the solid and vapor phases are not affected, since salt molecules will be separated
from water molecules in the freezing or evaporation process. For example, one can get fresh
water from melting the ice chunks in the arctic ocean as well as distilling sea water. In this
figure we can see that the intersection points with the lines of solid and vapor phases move in
opposite directions: the freezing point is lower, but the boiling point is higher.
If temperature remains the same and pressure can change, the expression of the Gibbs
function is modified into this formula:
dG D Vdp C dN: (5.17)
60 5. ADVANCED TOPICS

Solid
Solid

Chemical Potential

Chemical Potential
Liquid

Liquid
Vapor
Vapor

Temperature Pressure
(a) (b)

Figure 5.10: Chemical potential vs. (a) temperature and (b) pressure in different phases.

In the same way, we can find the slope of the function of the chemical potential vs. pres-
sure:    
@ @V
D > 0: (5.18)
@P T;n @N T;p

Figure 5.10b shows a sketch of chemical potential vs. pressure in different phases, and the
dashed line is the chemical potential of solution with salt added. Imagine drawing a horizontal
line in the liquid phase region; it will intersect these two parallel lines at different pressures. This
difference is called osmotic pressure when solutions with different concentrations of solute are
separated by a semipermeable membrane, which only allows the small water molecules to pass
through. This effect is shown in Fig. 5.11, where the difference in surface levels is caused by the
osmotic pressure.

Figure 5.11: Demonstration of osmotic pressure.


5.6. DISSIPATIVE STRUCTURES 61
Suppose the container in Fig. 5.11 is separately filled with fresh water and a saline solution
at the same level at first. Initially, the pressures in these two chambers are the same, but the
smaller water molecules in the fresh water chamber will cross the membrane and infiltrate into
the chamber of the saline solution. Gradually, the pressures in these two chambers will become
different, and the water level of the saline solution becomes higher than that of the fresh water. If
a turbine is connected between these two chambers, power can be generated from the difference
in pressure.
Most large rivers eventually flow into oceans, where the fresh water mixes with the salty
sea water. The difference in chemical potentials between them is just like the difference in tem-
perature, which can be tapped to generate power. In 2009, a pilot osmosis power plant was built
in Norway. It consumed 10 liters of fresh water and 20 liters of sea water per second, and it
could generate up to 4 kW of power. Although the supply of water is abundant, the cost of
the membrane was a limiting factor, so the effort of developing a larger scale power plant was
discontinued in 2013.
Osmotic pressure is a concept widely used in biology, since cell membrane is semiperme-
able. For example, if sufficient water is absorbed in a plant, the pressure in the cells gets rather
high, which provides the strength for the plant to stretch out its branches and leaves. In case of
severe drought, the pressure in the cells decreases, so the plant droops down. Unlike plant cells,
there is no wall in animal cells for protection, so it is dangerous if the internal pressure becomes
too high. In the extreme cases, animal cells can burst if too much water is absorbed. On the
other hand, if cells are immersed in solution with very high concentration of solute, water will
be sucked out of the cells and they will die. This is the principle of preserving food with salt,
where bacteria cannot survive such an environment. Similarly, if too much fertilizer is applied
to crops, they will wither due to loss of water.

5.6 DISSIPATIVE STRUCTURES


Looking around the world, we see two very different scenarios: there is destruction everywhere,
but at the same time there is also construction everywhere. For example, the pavement of road
breaks down and pit holes start to appear, buildings get old and eventually collapse, people age
and then pass away, even stars in the universe will eventually die. On the other hand, new roads
are paved, and new buildings are constructed, and babies and new stars are born all the time.
The second law of thermodynamics indicates that the arrow of time points toward the direction
of increasing entropy, which means the dissolving of forms or destruction. How do we explain
the spontaneous construction process, such as the appearance of life forms and evolution?
First, we need to make a distinction of two different kinds of systems: closed systems
and open systems. The behavior of closed systems, which do not exchange matter with the
environment, can be described by the principles discussed in the previous sections. However,
the behavior of open systems can be very different. Second, we also need to make a distinction
between equilibrium states and stable states, which can be far from equilibrium.
62 5. ADVANCED TOPICS
For example, in the summer and fall seasons, hurricanes are very common in the oceans.
The condition for hurricanes is a quasi-stable situation—far from equilibrium state: the warm
and humid air with lower density is under cool and dry air with higher density. Air masses with
different humidity do not mix easily, instead, the boundary between them is rather stable and
the convection process does not happen naturally. However, if a channel of convection is formed,
it becomes a short circuit between these two layers of air, and the surrounding air at the bottom
layer will funnel through it. Due to Coriolis force from the spin of the earth, the air will rotate
with very high speed. However, when a hurricane lands on continents, the supply of warm and
humid air disappears, and thus it cannot maintain its structure and will die out quickly. From
a certain point of view, a hurricane is similar to a life form, which depends on the supply of
nutrient and disposal of waste to survive.
Some closed systems can also demonstrate similar behaviors of life forms. For example,
when a viscous fluid in a pan is heated from the bottom, hurricane-shaped funnels will be
formed. If the viscosity of the fluid is high and the gradient of temperature is large, multiple
hurricane cells can show up all together. In addition, they will form a regular pattern as shown
in Fig. 5.12, which is called Bénard cells. In each cell, the hot fluid rises at the center, and heat
is released to the ambient, and then the cooled fluid falls at the rim. With the formation of such
a regular pattern, the heat transfer is more efficient.

T + ∆T

Figure 5.12: Bénard cells and convection pattern.

From a certain point of view, these systems work just like heat engines, which absorb
heat from a high-temperature source and release heat to a low-temperature source. In a broader
sense, high-quality energy is absorbed, and low-quality energy is dissipated. In this process, it
is possible to gain negative entropy so that organized structures can be formed and maintained.
Actually, all the life forms work in this way. For example, plants absorb sunlight—high-quality
energy—and release heat to the environment. In the meantime, carbon dioxide and water can
be synthesized into carbohydrates, and plants can survive and grow. In a similar way, animals
intake food and absorb the chemical energy in it, and then release heat and dispose waste to the
5.7. SOCIOECONOMIC SYSTEMS 63
environment. Without the negative entropy obtained in this way, life forms cannot maintain
their metabolic processes and will die.
All these systems we discussed above can be classified as dissipative structures, which have
two striking characteristics that are very different from the equilibrium states. First, they need
constant exchange of matter or energy with the environment. In this process, they consume
high-quality energy and dissipate low-quality energy. In this way, negative entropy is obtained.
Second, they can maintain steady dynamic structures with the first condition satisfied. In other
words, the negative entropy gained from the environment supports the formation of the steady
dynamics structures.

5.7 SOCIOECONOMIC SYSTEMS


When two people get married, it is similar to the situation of two atoms bonding together into
a molecule, which was discussed in Section 5.3. In this process, the entropy is reduced, since
the actions of these two people need to be correlated. In other words, each person has lost
some freedom in his/her actions. The benefit, from the point view of economy, is the reduced
living cost, because they can share the rent or mortgage. From this example we can see that the
terminology of thermodynamics can also be applied in socioeconomic systems.
• People ! Atoms or Molecules
• Money ! Energy
• Freedom ! Entropy
• Price ! (reversed) Temperature
In the pre-modern age, there was a clear division of labor at home. At that time, most
of the jobs in the society were not open to women, so men were the so-called “bread winners.”
However, women were also very busy at home. Besides cooking meals and looking after children,
they also needed to make the clothes and shoes for the whole family, as well as washing them
by hand. At that time, the relationship of a couple was complementary, just like an ionic bond
of NaCl.
Now women are liberated from these chores at home, and almost all the careers are open
to them. Therefore, the relationship of a couple is cooperative, just like a covalent bond of O2 .
Unlike in ancient times, love plays a much more important role in modern marriage. However,
economic factor still has some influence. Data shows that there is a slight drop in divorce rate
during recessions, since the extra cost of living separately prohibits some couples from divorcing.
If people are modeled as atoms, then the military units are just like crystals. This is demon-
strated clearly during parades, where blocks of soldiers march forward with locked steps. In an-
cient times, a battle between two armies was very similar to the collision between two rocks.
The army with a weaker bond and lack of discipline will be defeated, just like the softer rock is
smashed by the harder one in the collision.
64 5. ADVANCED TOPICS
There are also heat engines in the society, and they are the merchants. Their business model
is very simple: purchase commodities at lower prices (high temperature) and sell them at higher
prices (low temperature). The profit a merchant makes is just like the work obtained from a heat
engine.
Enterprises are like animals, which absorb the energy from the food and then use the
energy to obtain more food. In the same way, enterprises maintain their operation by constantly
generating revenues, which is used to pay the salary and supplies, etc. If their business model
is very profitable, they can grow in scale, just like the growth of young animals when abundant
food is available. On the other hand, if there is a trouble in cash flow, such a cycle will wind
down and the enterprises might go to bankruptcy.
A society, which comes in various forms, is an organization of people. Ancient agricul-
tural societies were loosely organized, and each family was able to live a self-sufficient life. Such
societies are just like an ideal gas, where the interaction between the molecules is very weak.
Another analogy for such a society is sand—tightly bound within each family but with very
weak connections at higher levels. Socialist society with planned economy is the other extreme,
and every citizen is bonded into a single large organization. An analogy of such societies is a
mammoth machine, where all the components are linked together. Such a society can be very
efficient in war and industry production, but it is not able to evolve and eventually will dete-
riorate. Societies with market economy is like an ecosystem—various organizations form food
chains and become mutually dependent. Such societies are quite dynamic, new enterprises are
constantly created while unsuccessful old ones become acquired or demised. In addition, the
whole economy has periods of expansion and contraction. More importantly, this type of soci-
ety can evolve with the advancement of technology, therefore, it is the most successful form of
organization.
65

Bibliography
[1] Claus Borgnakke and Richard E. Sonntag, Fundamentals of Thermodynamics, 8th ed., Wi-
ley, 2012.

[2] Michael Moran and Howard Shapiro, Fundamentals of Engineering Thermodynamics, 8th
ed., Wiley, 2014.

[3] Yunus Cengel and Michael Boles, Thermodynamics: An Engineering Approach, 8th ed.,
McGraw-Hill, 2014.

[4] Stephen Blundell and Katherine Blundell, Concepts in Thermal Physics, 2nd ed., Oxford
University Press, 2009. DOI: 10.1093/acprof:oso/9780199562091.001.0001.

[5] Carl Helrich, Modern Thermodynamics with Statistical Mechanics, Springer, 2009. DOI:
10.1007/978-3-540-85418-0.

[6] Robert Swendsen, An Introduction to Statistical Mechanics and Thermodynamics, Oxford,


2012. DOI: 10.1093/acprof:oso/9780199646944.001.0001.

[7] Michael Starzak, Energy and Entropy: Equilibrium to Stationary States, Springer, 2010.
DOI: 10.1007/978-0-387-77823-5.

[8] J. S. Dugdale, Entropy and its Physical Meaning, 2nd ed., Taylor & Francis, 1996. DOI:
10.4324/9780203211298.

[9] Arieh Ben-Naim, Entropy and the Second Law: Interpretation and Miss-Interpretations,
World Scientific Publishing Company, 2012. DOI: 10.1142/8333.

[10] Robert Holyst and Andrzej Poniewierski, Thermodynamics for Chemists, Physicists, and En-
gineers, Springer, 2012. DOI: 10.1007/978-94-007-2999-5.

[11] J. M. Smith, H. C. Van Ness, and M. Abbott, Introduction to Chemical Engineering Ther-
modynamics, 7th ed., McGraw-Hill, 2004. DOI: 10.1021/ed027p584.3.

[12] Dilip Kondepudi and Ilya Prigogine, Modern Thermodynamics: From Heat Engines to Dis-
sipative Structures, Wiley, 2014. DOI: 10.1002/9781118698723.

[13] Ilya Prigogine and Isabelle Stengers, Order Out of Chaos, Bantam, 1984. DOI:
10.1063/1.2813716.
66 BIBLIOGRAPHY
[14] James Stone, Information Theory: A Tutorial Introduction, Sebtel Press, 2015. DOI:
10.13140/2.1.1633.8240.
[15] Thomas Cover and Joy Thomas, Elements of Information Theory, 2nd ed., Wiley 2006.
DOI: 10.1002/0471200611.
[16] Arieh Ben-Naim, Molecular Theory of Water and Aqueous Solutions: Understanding Water,
World Scientific Publishing Company, 2009. DOI: 10.1142/9789812837615.
[17] Robert Ayres, Information, Entropy, and Progress: A New Evolutionary Paradigm, Ameri-
can Institute of Physics, 1994.
[18] Reiner Kümmel, The Second Law of Economics: Energy, Entropy, and the Origins of Wealth,
Springer, 2011. DOI: 10.1007/978-1-4419-9365-6.
67

Author’s Biography
YUMIN ZHANG
Yumin Zhang is a professor in the Department of Engineering & Technology, Southeast Mis-
souri State University. In 1981, he was admitted into the Department of Engineering Mechan-
ics, Tsinghua University in China, and his career goal was in the field of aerospace engineering.
However, his dream was shattered by the course of thermodynamics, since he had great trou-
ble learning it. His experience was not an exception; generations of engineering students have
studied thermodynamics in college and many of them felt confused by the subject matter.
In fall 2008, Yumin received a teaching assignment in a thermodynamics course and there-
fore had to relearn it. After a quarter of a century, he had become more mature from his experi-
ence in learning and research and therefore he found that thermodynamics was not hard at all; in-
stead, it was very beautiful. Survey results from his students in 2016 showed that this course was
not more difficult than other engineering courses, such as Engineering Mechanics—Dynamics.
Therefore, he would like to share his understanding and approaches with more students and
hopes that this book can help them overcome the challenges of learning thermodynamics.

You might also like