You are on page 1of 609

CHEMICAL

ENGINEERING
THERMODYNAMICS - I
SI UNITS

FOR
DEGREE COURSE IN CHEMICAL ENGINEERING OF ALL UNIVERSITIES IN INDIA

K. A. GAVHANE
Formerly, Vice-Principal & Head of Chemical Engg. Dept.,
S.E. Society's Satara Polytechnic,
Satara.

N1075
CHEMICAL ENGINEERING THERMODYNAMICS – I ISBN – 978-93-80064-66-6
Sixth Edition : December 2017
© : Author
The text of this publication, or any part thereof, should not be reproduced or transmitted in any form or stored in any computer storage
system or device for distribution including photocopy, recording, taping or information retrieval system or reproduced on any disc, tape,
perforated media or other information storage device etc., without the written permission of Author with whom the rights are reserved.
Breach of this condition is liable for legal action.
Every effort has been made to avoid errors or omissions in this publication. In spite of this, errors may have crept in. Any mistake, error
or discrepancy so noted and shall be brought to our notice shall be taken care of in the next edition. It is notified that neither the publisher
nor the author or seller shall be responsible for any damage or loss of action to any one, of any kind, in any manner, therefrom.

Published By : Polyplate
NIRALI PRAKASHAN
Abhyudaya Pragati, 1312, Shivaji Nagar,
Off J.M. Road, PUNE – 411005
Tel - (020) 25512336/37/39, Fax - (020) 25511379
Email : niralipune@pragationline.com

☞ DISTRIBUTION CENTRES
PUNE
Nirali Prakashan : 119, Budhwar Peth, Jogeshwari Mandir Lane, Pune 411002, Maharashtra
Tel : (020) 2445 2044, 66022708, Fax : (020) 2445 1538
Email : bookorder@pragationline.com, niralilocal@pragationline.com
Nirali Prakashan : S. No. 28/27, Dhyari, Near Pari Company, Pune 411041
Tel : (020) 24690204 Fax : (020) 24690316
Email : dhyari@pragationline.com, bookorder@pragationline.com
MUMBAI
Nirali Prakashan : 385, S.V.P. Road, Rasdhara Co-op. Hsg. Society Ltd.,
Girgaum, Mumbai 400004, Maharashtra
Tel : (022) 2385 6339 / 2386 9976, Fax : (022) 2386 9976
Email : niralimumbai@pragationline.com

☞ DISTRIBUTION BRANCHES
JALGAON
Nirali Prakashan : 34, V. V. Golani Market, Navi Peth, Jalgaon 425001,
Maharashtra, Tel : (0257) 222 0395, Mob : 94234 91860
KOLHAPUR
Nirali Prakashan : New Mahadvar Road, Kedar Plaza, 1st Floor Opp. IDBI Bank
Kolhapur 416 012, Maharashtra. Mob : 9850046155
NAGPUR
Pratibha Book Distributors : Above Maratha Mandir, Shop No. 3, First Floor,
Rani Jhanshi Square, Sitabuldi, Nagpur 440012, Maharashtra
Tel : (0712) 254 7129
DELHI
Nirali Prakashan : 4593/21, Basement, Aggarwal Lane 15, Ansari Road, Daryaganj
Near Times of India Building, New Delhi 110002, Mob : 08505972553
BENGALURU
Pragati Book House : House No. 1, Sanjeevappa Lane, Avenue Road Cross,
Opp. Rice Church, Bengaluru – 560002.
Tel : (080) 64513344, 64513355,Mob : 9880582331, 9845021552
Email:bharatsavla@yahoo.com
CHENNAI
Pragati Books : 9/1, Montieth Road, Behind Taas Mahal, Egmore,
Chennai 600008 Tamil Nadu, Tel : (044) 6518 3535,
Mob : 94440 01782 / 98450 21552 / 98805 82331,
Email : bharatsavla@yahoo.com
niralipune@pragationline.com | www.pragationline.com
Also find us on www.facebook.com/niralibooks
Note : Every possible effort has been made to avoid errors or omissions in this book. In spite this, errors may have crept in. Any type of error or
mistake so noted, and shall be brought to our notice, shall be taken care of in the next edition. It is notified that neither the publisher, nor the author
or book seller shall be responsible for any damage or loss of action to any one of any kind, in any manner, therefrom. The reader must cross check
all the facts and contents with original Government notification or publications.
ek>kh vkbZ

DSQ= jQfDe.khckbZ vac* "kh xOgk.ks


o

ek>ks oMhy

DSQ= vac* "kh egknso xOgk.ks


;kaps pj.kh lefiZr
PREFACE TO THE SIXTH EDITION

It gives me a great pleasure to present the sixth edition of the book -


"Chemical Engineering Thermodynamics-I" to the students of degree course
in Chemical Engineering of all Universities in India.

The subject matter is presented in a lucid and simple language with neat
illustrations along with a fairly large number of solved examples.

I hope that this book would fill a long felt need for a comprehensive book on
Chemical Engineering Thermodynamics.

I acknowledge the constant encouragement from my colleagues, friends and


family members.

I am very thankful to the publishers, Shri. Dineshbhai Furia, Jigneshbhai


Furia and the entire staff members of Nirali Prakashan, Pune for bringing out the
book neatly in a short time.

I hope that students and teachers will find the matter presented in the book
very useful.

I would welcome and appreciate suggestions and comments from students


and teachers for improving the quality of the book.

K. A. GAVHANE

Mob. No. 9850242440

kagavhane@yahoo.in
PREFACE TO THE FIRST EDITION

It gives me a great pleasure to present the first edition of the book


– "Chemical Engineering Thermodynamics-I" to students of degree course in
Chemical Engineering of all Universities in India.

The subject matter is presented in a lucid and simple language with neat
illustrations along with a fairly large number of solved examples.

I hope that this book would fill a long felt need for a comprehensive book of
Chemical Engineering Thermodynamics.

I acknowledge the constant encouragement from may colleagues, friends and


family members.

I am very thankful to the publishers, Shri. Dineshbhai Furia, Jigneshbhai


Furia and the entire staff members of Nirali Prakashan, Pune for bringing out the
book neatly in a short time.

I hope that students and teachers will find the matter presented in the book
very useful.

I would welcome and appreciate suggestions and comments from students


for improving the quality of the book.

K. A. GAVHANE

Mob. No. 9850242440

kagavhane@yahoo.in
CONTENTS
1. Introduction and Basic Concepts 1.1 – 1.58
• Thermodynamics - definitions 1.1
• Brief discussion regarding laws of thermodynamics 1.1
• Utility of thermodynamics 1.2
• Systems of units 1.3
• System and surroundings 1.6
• Phase 1.7
• Homogeneous system 1.8
• Heterogeneous system 1.8
• Types of thermodynamic system 1.8
• Steady state 1.9
• Thermodynamic properties of system 1.10
• State of a system 1.11
• State functions 1.12
• Path functions 1.13
• Process 1.14
• Types of process 1.14
• Energy 1.15
• Thermodynamic equilibrium 1.17
• Reversible and irreversible processes 1.18
• Reversible work of expansion 1.21
• Heat 1.25
• Temperature and zeroth law of thermodynamics 1.26
• Ideal gas temperature scale 1.27
• IPT scale 1.29
• Preliminary mathematical background 1.30

• Solved Examples 18 1.40 to 1.58

2. First Law of Thermodynamics 2.1 – 2.152


• Statements of first law of thermodynamics 2.1
• First law for a closed system undergoing a cycle 2.2
• First law for a closed system undergoing a change of state 2.3
• Energy - a property of the system 2.5
• Energy of an isolated system 2.7
• PMM-1 2.7
• Enthalpy of a system 2.9
• Heat capacity 2.10
• Molar heat capacity at constant pressure 2.10
• Molar heat capacity at constant volume 2.11
• Reversible constant volume process 2.12
• Reversible constant pressure process 2.13
• Reversible constant temperature process 2.15
• Reversible adiabatic process 2.16
• Polytropic process 2.23
• First law for steady state flow process 2.28
• Compressors 2.35
• Turbines 2.41
• Pumps 2.43
• Nozzles 2.44
• Throttling process 2.46
• Joule - Thomson coefficient 2.50

• Solved Examples 70 2.53 – 2.152

3. P-V-T Behaviour and Heat Effects 3.1 – 3.96


• Graphical representation of P-V-T behaviour 3.1
• P-V diagram 3.1
• P-T diagram/phase diagram 3.4
• Mathematical representation of P-V-T behaviour 3.5
• Ideal gas law 3.6
• Equation of state for real gases 3.8
• van der Waals equation of state 3.11
• Principle of corresponding states 3.14
• Compressibility charts 3.16
• Virial equation of state 3.19
• Redlich-Kwong equation 3.21
• Soave-Redlich-Kwong equation of state 3.21
• Peng-Robinson (PR) equation of state 3.22
• Heat effects 3.22
• Standard states 3.23
• Heat of reaction 3.23
• Standard heat of reaction 3.23
• Standard heat of formation 3.24
• Standard heat of combustion 3.24
• Standard heat of reaction from standard heats of formation 3.26
• Standard heat of reaction from standard heats of combustion 3.27
• Hess's law of constant heat summation 3.28
• Effect of temperature on the standard heat of reaction 3.30
• Adiabatic flame temperature 3.32

• Solved Examples 33 3.35 to 3.96

4. Second Law of Thermodynamics 4.1 to 4.106


• Utility of second law of thermodynamics 4.1
• Statements of second law of thermodynamics 4.1
• Heat engine 4.2
• Working medium 4.2
• Thermal reservoirs 4.2
• Source and sink 4.3
• Refrigerator and heat pump 4.5
• Refrigerator 4.5
• Heat pump 4.7
• PMM-2 4.8
• Possible and impossible engine as per Kelvin-Planck statement 4.9
• Possible and impossible system as per Clausius statement 4.10
• Equivalence of Kelvin-Planck and Clausius statements 4.10
• Carnot cycle 4.12
• Carnot theorems 4.16
• Maximum efficiency of heat engine 4.17
• Thermodynamic temperature scale 4.17
• COP of a Carnot refrigerator/heat pump 4.19
• Two reversible adiabatic paths cannot intersect each other 4.20
• Reversible adiabatic lines are steeper than isothermal lines 4.21
• Any reversible process can be replaced by two reversible and one
isothermal process 4.22
• Clausius theorem and Clausius inequality 4.23
• Entropy 4.26
• Characteristics of entropy 4.29
• Units of entropy 4.29
• Physical significance of entropy 4.30
• Principle of entropy increase 4.30
• Applications of principle of increase of entropy 4.32
• Heat transfer through a finite temperature difference 4.32
• Adiabatic mixing of two fluids 4.33
• Isothermal mixing of ideal gases 4.34
• Temperature-entropy plot 4.35
• Entropy changes for an ideal gas 4.38
• Entropy change during phase change 4.40
• Third law of thermodynamics 4.41
• Entropy change in chemical reactions 4.42
• Entropy generation in a closed system 4.43
• Entropy generation in an open system 4.44
• Isentropic efficiency 4.44
• Combined first and second law 4.45
• Comparison of first law with second law of thermodynamics 4.46

• Solved Examples 55 4.47 to 4.106

5. Thermodynamic Properties of Pure Fluids 5.1 to 5.140


• Thermodynamic properties 5.1
• Helmholtz free energy 5.1
• Gibbs free energy 5.2
• Fundamental property relations 5.3
• Maxwell's equations 5.5
• Clapeyron equation 5.6
• Clausius - Clapeyron equation 5.8
• Vapour pressure-temperature relationship 5.9
• Volume expansivity and isothermal compressibility 5.9
• Entropy-heat capacity relationships 5.10
• Difference between heat capacities 5.11
• Ratio of heat capacities 5.14
• Differential equation for entropy 5.16
• Differential equation for internal energy 5.20
• Differential equations for enthalpy 5.22
• Effect of T, P and V on U, H and S 5.23
• Joule-Thomson coefficient 5.24
• Gibbs-Helmholtz equation 5.25
• Jacobian method 525
• Fugacity 5.27
• Fugacity coefficient 5.28
• Activity 5.28
• Choice of standard state 5.29
• Effect of T and P on fugacity 5.29
• Determination of fugacity of real gases 5.30
• Residual volume method 5.30
• Approximate method 5.32
• Generalised method 5.33
• Determination of fugacity from equation of state 5.34
• Fugacities of pure solids and liquids 5.35
• Departure functions 5.36
• Thermodynamic diagrams 5.40
• P-H diagram 5.40
• H-T diagram 5.41
• T-S diagram 5.42
• H-S diagram 5.42
• Solved Examples 59 5.43 to 5.140
6. Refrigeration 6.1 – 6.44
• Refrigeration - definitions 6.1
• Applications of refrigeration 6.1
• Types of refrigerators 6.2
• Refrigerating effect and refrigeration unit capacity 6.2
• Reversed Carnot cycle 6.3
• Refrigeration systems 6.4
• Air refrigeration system 6.5
• Vapour compression refrigeration system 6.10
• Analysis of vapour compression system 6.12
• Advantages and disadvantages of vapour compression system over air
refrigeration system 6.14
• Vapour absorption refrigeration system 6.14
• Comparison between vapour compression and vapour absorption system 6.17
• Refrigerants 6.17
• Desirable properties of refrigerants 6.17
• Linde process 6.21
• Claude process 6.23
• Solved Examples 12 6.24 to 6.44
❐❐❐
Chapter ... 1
INTRODUCTION AND BASIC CONCEPTS

• Thermodynamics is the branch of science which deals with all forms of energy and the
interconversion of different forms of energy.
• Thermodynamics is the branch of science which is concerned with the relations between
heat and other forms of energy involved in physical and chemical processes.
• Two forms of energy, namely heat and work are generally associated with changes
occurring in the universe. Thermodynamics is mainly a study of heat and work effects.
• Thermodynamics is the branch of science which deals with transformations of energy of
all kinds from one form to another and the laws governing such transformations.
• The general limitations within which transformations of energy of all kinds, from one
form to another, are observed to occur are known as the first and second laws of
thermodynamics.
• In chemical engineering thermodynamics, more emphasis (stress/importance) is given to
the treatment of properties of solutions, phase equilibria and chemical reaction
equilibria instead of to the thermodynamic analysis of heat engines and heat power
cycles.
• Thermodynamics deal with macroscopic systems. Macroscopic systems (those
consisting of a large number of molecules, e.g., a gas mixture containing CH4, O2 and
N2) are the basis of human observation and experience. Therefore, no exception to the
laws of thermodynamics has yet been observed.
• In the study of thermodynamics, we are basically concerned with the four laws : the
zeroth, first, second and third laws of thermodynamics. The laws of thermodynamics
apply to matter in bulk (macroscopic systems) and not to individual atoms or molecules.
The laws deal with the macroscopic changes of interest in the problem handled and
because of this the study of thermodynamics is of direct benefit in the analysis of
industrial problems. These laws are based on a large number of experimental
observations and logical reasoning like all physical laws. The laws of thermodynamics
have no mathematical proof (i.e., these laws cannot be proved as we prove mathematical
theorems) and have not so far/yet been violated (no experimental evidence is available
so far to doubt the validity of these laws). The laws of thermodynamics are merely a
(1.1)
Chemical Engineering Thermodynamics - I 1.2 Introduction and Basic Concepts

systematization of observations of common human experience. Several direct and


indirect consequences of these laws may be deduced strictly on logical grounds with the
application of a moderate knowledge of mathematics.
The zeroth law of thermodynamics deals with thermal equilibrium between a set of bodies
and forms the basis for the measurement of temperature. The first law of thermodynamics
indicates that there is an exact equivalence between the various forms of energy involved in any
process. The second law of thermodynamics provides a means of predicting whether a given
thermodynamic process is possible or not. It helps us to predict whether a given process is
feasible or not. The first law of thermodynamics leads to the concept of internal energy, while
the second law of thermodynamics leads to the concept of entropy. The third law of
thermodynamics places a limitation upon the value of entropy. The law defines the absolute zero
of entropy and helps us to calculate the absolute entropies of pure substances. The zeroth law of
thermodynamics was formulated by Fowler in the year 1931. However, since the first and second
laws of thermodynamics were already existed at that time, it was named as zeroth law of
thermodynamics so that it proceeds the first and second laws to form a logical sequence. In case
of internal energy and entropy functions, we are interested to know the changes in these
functions. Actually, the first and second laws of thermodynamics do not give any idea/clue as to
the method of finding the absolute values of the internal energy and entropy.
UTILITY OF THERMODYNAMICS :
Thermodynamics tells us whether a given process, including a chemical reaction, is possible
/ feasible or not, under a given set of conditions (e.g., temperature, etc.). However, it does not tell
anything about the rate at which a given process will take place / proceed. That is,
thermodynamics cannot predict whether a specified change will occur in an acceptable period of
time or not. Thermodynamics provides information regarding how far a given process may
proceed (the maximum possible extent to which a given process can proceed) and the degree to
which the energy present in a system can be utilized to cause a given process to occur / take
place. It helps to make a preliminary choice of the process variables, their effect on the extent of
a process and to find out thermal / heat effects associated / involved in a given process.
As far as chemical reactions are concerned, the aspects such as whether certain reactions can
take place, to what extent chemical conversions can occur, the effect of parameters like
temperature and pressure on the behaviour of a chemical reaction, the composition of product
mixtures if equilibrium is reached and the amount of heat released / absorbed can be determined
from thermodynamic calculations alone. However, thermodynamics cannot tell anything about
the speed / rate of the reaction and the effect of the shape of the reactor used (kinetics provides
this). A knowledge of both thermodynamics and kinetics is essential for a complete
understanding of the unit processes. Thermodynamics of a chemical reaction must be known
before its kinetics can be utilized. Before knowing how fast a chemical reaction will go / take
place, one must know can it go at all / can it take place at all and, if so, how far (to what extent).
Thermodynamics answers this.
Chemical Engineering Thermodynamics - I 1.3 Introduction and Basic Concepts

Thermodynamics enables us to calculate the maximum efficiency of a turbine or a


refrigerator and the maximum achievable yield in a given chemical reaction.

INTERNATIONAL SYSTEM OF UNITS

In the study of thermodynamics we deal with a large number of physical quantities such as
temperature, pressure, volume, energy, etc. These are measurable in terms of numbers and units
of measurements. A unit is an arbitrarily selected standard of measure for a physical quantity.
Therefore, volume may be measured in litres, cubic meters, etc. There are four systems of units -
SI, MKS, CGS and FPS. The unit system of this book is the International System of Units,
abbreviated as SI units. A thorough knowledge of the units of commonly encountered physical
quantities is essential in the understanding of thermodynamics.

Physical quantities are divided into two groups as primary / fundamental quantities and
secondary / derived quantities. Primary quantities such as mass, length, time and temperature are
called dimensions or base units and are represented by symbols M, L, θ and T. Secondary
quantities are derived from the fundamental quantities, e.g., velocity, which involve length and
time; density, which involves mass and length. These are measured in terms of the derived units -
the units obtained by multiplying and / or dividing the base units.

BASIC SI UNITS

Mass : kilogram, kg

Length : meter, m

Time : second, s

Temperature : kelvin, K

Force : newton, N

Pressure : newton/(meter)2, N/m2

Energy : newton.meter, N.m = joule, J

Power : (newton-meter)/second, (N.m)/s = J/s = watt, W

The first letter in the abbreviated form of the SI units named after scientists should be written
in upper case and all other units/unit symbols should be written in lower case letters.
For example, meter ⇒ m, second ⇒ s, pascal ⇒ Pa, newton ⇒ N, kelvin ⇒ K.

PREFIXES FOR SI UNITS

The SI prefixes are used in order to express units in terms of the multiples of SI units.
Multiples in power of 10 are formed with certain prefixes, which are summarized in Table 1.1
when the size of units is too large or too small.
Chemical Engineering Thermodynamics - I 1.4 Introduction and Basic Concepts

Table 1.1 : SI Prefixes


Amount Multiple / Factor Prefix Symbol
10 101 deka da
100 102 hecto h
1000 103 kilo k
1000 000 106 mega M
1000 000 000 109 giga G
0.1 10–1 deci d
0.01 10–2 centi c
0.001 10–3 milli m
0.000 0001 10–6 micro µ
0.000 000 0001 10–9 nano n
For example : 1000000 N = 1 MN
Thus, the standard atmospheric pressure (101325 Pa) can be expressed as 101.325 kPa or
0.101325 MPa = 0.10 MPa.
FORCE
The SI unit of force is newton (N), which is derived from Newton’s law of motion. This law
states that force is proportional to the product of mass and acceleration.
F ∝ m· a
∴ F = k ma
In the SI system, the constant k is unity and therefore
F = m·a
The newton (N) is defined as the force which when applied to a body of mass 1 kg gives it an
acceleration of 1 m/s2. Hence, the newton is a derived unit representing 1 (kg·m)/s2. When a body
of mass 1 kg is accelerated by 1 m/s2, then the force acting on the body is 1 (kg·m)/s2 and is
designated as 1 newton (1 N).
1 N = 1 (kg·m)/s2
The unit of force newton, abbreviated as N, has been named after the scientist Newton.
PRESSURE
Pressure of a fluid on a surface is defined as the normal force exerted by the fluid on the unit
area of the surface. The SI unit of pressure derived from this definition is thus newton per square
meter (N/m2). The unit of pressure N/m2 has been named as pascal, abbreviated as Pa, in honour
of the scientist Pascal.
Chemical Engineering Thermodynamics - I 1.5 Introduction and Basic Concepts

The MKS unit of pressure is kilogram-force per square centimeter, abbreviated as kgf/cm2.
At the mean sea level, the air of the atmosphere exerts a pressure of 101325 Pa on all bodies
exposed to it and this pressure is known as the standard or normal atmospheric pressure, 1 atm.
One atmospheric pressure (1 atm) balances a column of 760 mm Hg at 0oC.
1 atm = 760 mm Hg
= 101325 N/m2 (Pa)
= 101.325 kPa
= 1.103 bar
= 1.033 kgf/cm2
= 14.7 psi (lbf/in2)
= 10.33 m H2O
= 760 torr
Sub-atmospheric pressure (vacuum), i.e., pressure below the actual atmospheric pressure is
expressed in torr (in honour of the scientist Torricelli).
Usually, pressure is measured with the help of a pressure gauge. The gauge registers /
records the difference between the pressure prevailing in the vessel on which it is fixed and the
local atmospheric pressure. The pressure indicated by the pressure gauge is called the gauge
pressure and therefore, the letter g follows the unit of pressure, e.g., 30 psig, 5 (kgf/cm2) g.
To obtain the absolute / true pressure prevailing in the vessel, we have to add the local
atmospheric or barometric pressure to the gauge pressure, both are to be expressed in consistent
units.
The relationship between absolute pressure and gauge pressure is
Absolute pressure = Gauge pressure + Atmospheric pressure
The relationship between absolute pressure and vacuum is
Absolute pressure = Atmospheric pressure – Vacuum
Absolute pressures must be used in thermodynamic calculations.
Since the unit pascal (Pa) is very small in magnitude, for convenience, the pressure is usually
expressed in kilopascal (kPa) and megapascal (MPa). The bar, which is a multiple of pascal,
is also used as a unit of pressure in the SI system. It is approximately / roughly equal to one
atmospheric pressure (1 atm).
1 bar = 105 Pa
WORK / ENERGY AND POWER
Work is defined as the product of the force acting on a body and the distance travelled by the
body in the direction of the applied force.
W = F × L ⇒ newton × meter ⇒ N·m = joule
Chemical Engineering Thermodynamics - I 1.6 Introduction and Basic Concepts

The SI unit of work is thus newton.meter (N·m) and has been given a special name joule,
abbreviated as J.
1 N·m = 1 J = 1 (kg·m2)/s2
Since the energy has the same dimensions as work [ML2 θ–2], the unit of energy (whether
mechanical, thermal or electrical) in the SI system is joule (J).
Power is defined as the work done per unit time.
P = W/θ
In the SI system, power is measured in joules per second (J/s) and this has been given a
special name watt, abbreviated as W.
1 W = 1 J/s = 1 (N·m)/s = 1 (kg·m2)/s2
HEAT
It is energy in transit. It is a form of energy that passes from one body to another, either by a
direct contact or by means of radiation, solely as the result of a difference in temperature.
Heat, like work, is measured in joules. The unit of heat in the SI system is joule (J). Other
units used for heat are the calorie (cal) and the British thermal unit (Btu).
1 cal = 4.1868 J
1 Btu = 1055.056 J
BASIC CONCEPTS AND DEFINITIONS
It is very essential to understand the meaning of certain terms which are associated with the
development of the laws of thermodynamics and their consequences. A clear perception
(understanding) of these terms results in achieving a full appreciation of these laws.
SYSTEM AND SURROUNDINGS
The specified portion of the universe containing a definite quantity/amount of a specific
substance or group of substances under study for the purpose of thermodynamic analysis of a
problem is called a system (thermodynamic system).
The remaining part of the universe other than the system is called the surroundings -
everything external to the system is called the surroundings. Usually the region of the universe in
the immediate vicinity of the system and getting affected by changes occurring in the system is
taken / considered as its surroundings. For example, when the contents of a jacketed reactor are
considered as the system, the steam condensing in the jacket may be treated as the surroundings.
It may be a water-bath in which the system under examination is immersed.
A system may be as simple as a certain amount of liquid contained in a vessel or it can be as
complex as a petrochemical plant or thermal power plant. A gas enclosed in a cylinder fitted with
Chemical Engineering Thermodynamics - I 1.7 Introduction and Basic Concepts

a piston, a vessel containing liquid, a distillation column, a reaction vessel are some examples of
system.
The combination of a system and its surroundings is called the universe.
A system may be separated from its surroundings by a real or imaginary boundary through
which energy (either as heat or some form of work) and / or mass may pass. The boundary may
be either fixed or moving.
The walls or boundaries of a system may be adiabatic or diathermal. The adiabatic walls of a
system do not permit the influence of its surroundings to change the state of the system contained
by them, e.g., double evacuated walls of a thermos and the walls of a refrigerator are examples of
adiabatic walls. On the other hand, the diathermal walls of a system permit the influence of its
surroundings to change the state of the system contained by them, e.g., walls of metals, glass, etc.
are examples of diathermal (non-adiabatic) walls. (Diathermal means which permits heat to pass
/ flow - a wall which allows the flow of heat through it is called a diathermal wall.)
A vessel containing liquid forms a system with the walls of the vessel forming a real
(system) boundary. The boundary of a system may change in shape as well as in size during a
given process. For example, if we treat a gas contained in a piston-cylinder assembly as our
system and if this is brought in contact with a hot body, the temperature as well as the volume of
the gas increases - one boundary surface moves as the gas expands (due to rise in temperature)
and the volume increases. In the case of an elastic spherical balloon containing a certain amount
of gas the system boundary can be changed by doing work - the spherical gas balloon can be
deformed into, say, a cylindrical shape by doing work.
PHASE
A pure substance is one which contains only one chemical species. Ice, water and steam have
the same chemical species. So ice, water, steam, a system consisting of ice, water and steam,
steam - water mixture are regarded as pure substances.
Three states of physical aggregation of matter are solid, liquid and gas. The different states in
which a substance (matter) exists are called phases (liquid phase, gas phase, solid phase).
A quantity of matter which is uniform throughout in physical state and chemical composition
is called a phase, e.g., water is one phase, but ice is another phase. For uniformity / homogeneity
of chemical composition, it is not necessary that the phase contains only a single chemical
species. A mixture of gas or a solution is uniform in chemical composition and therefore,
constitutes / forms a phase. Uniformity of physical state means that the material is all solid,
all liquid or all gas. Therefore, a gas mixture is one phase and a system containing water and
water vapour together contains two phases, a liquid and a gas phase. A system containing two
miscible liquids (e.g., water and methyl alcohol or ethyl alcohol) forms a single liquid phase
whereas a system containing two immiscible liquids (e.g., water and benzene) forms two liquid
phases.
Chemical Engineering Thermodynamics - I 1.8 Introduction and Basic Concepts

HOMOGENEOUS SYSTEM
A system consisting of a single phase is called a homogeneous system. Or a system which is
completely uniform throughout in physical state and chemical composition is called a
homogeneous system. (A system consisting of a gas, or a mixture of gases, or a pure liquid,
or solid or a solution ⇒ a mixture of air and water vapours, dilute nitric acid, a solution of
ammonia in water, a solution of sugar in water, i.e., a sugar solution, liquid water in a beaker are
some examples of homogeneous system.)
HETEROGENEOUS SYSTEM
A system consisting of more than one phase (two or more phases) is called a heterogeneous
system. The phases are separated from one another by definite boundary surfaces. This system is
not uniform throughout in its physical state.
A system consisting of a liquid and its vapour or of two immiscible liquids ⇒ water and
water vapour in a closed container, a liquid mixture of water and benzene or water and oil are
some examples of heterogeneous system.
TYPES OF SYSTEM
Closed System :
• A system of fixed mass is called a closed system.
• A system which can exchange energy but not matter with its surroundings is called a
closed system.
If the mass within the boundary of the system remains constant, it is called a closed system.
Therefore, a closed system does not permit the transfer of mass across its boundaries but permits
the transfer of energy and its mass remains constant. A batch reactor, a gas enclosed in a cylinder
fitted with a piston and a thermal power plant consisting of boiler, turbine, condenser and pump
are examples of closed system.
OPEN SYSTEM
• A system in which matter (and energy also) crosses the boundary of the system is called
an open system.
• A system which can exchange matter as well as energy with its surroundings is called an
open system.
• If both the mass and energy cross the boundaries of a system, it is called an open
system.
An open system permits transfer of both mass and energy across its boundaries and the mass
within the system may not remain constant. This system is also referred to as a flow system
wherein material flows into and out of the system. A tubular flow reactor, an air compressor and
a heat exchanger are examples of open system.
Chemical Engineering Thermodynamics - I 1.9 Introduction and Basic Concepts

ISOLATED SYSTEM
A system which can exchange neither matter nor energy through the boundaries with its
surroundings is called an isolated system. It is of fixed mass and energy. This is a system which
is not affected / influenced by the changes in its surroundings or environment. A perfectly
isolated system is an ideal concept that cannot be observed or attained in practice. The
combination of a system and its surroundings constitutes an isolated system. Therefore, the
universe can be treated as an isolated system.

ADIABATIC SYSTEM
A system which is thermally insulated from its surroundings is called an adiabatic system.
This system does not give heat to the surroundings nor does it receive heat from the
surroundings. In case of this system, there is no heat interaction of the system with its
surroundings.
Surroundings Surroundings

Boundary
Boundary
System Energy System
in Energy out

(a) A thermodynamic system (b) A closed system


Surroundings Surroundings
Boundary
Mass
in Mass out
System
Energy
Energy System
in
out

(c) An open system (d) An isolated system


Fig. 1.1 : System and its types
STEADY STATE

A steady state is defined as one in which the macroscopic properties (conditions) such as
temperature, pressure, etc. at a specified location in the system do not vary / change with time.
The conditions may be changing / varying from location to location within the system, but at any
one location they do not change / vary with time (i.e., the properties / conditions at all locations
in the system are constant/invariant with time).
Chemical Engineering Thermodynamics - I 1.10 Introduction and Basic Concepts

A system is said to be in a steady state if the macroscopic properties at a specified location in


the system do not change / vary with time. Consider a plane wall, the inside of which is at T1 and
the outside is at T2 such that T1 > T2. Since T1 > T2, heat will flow from the inside to the outside
and the temperature at a specified location in the wall varies with time. Ultimately, the wall
attains a steady state with respect to heat transfer and due to this the temperature at any given
location in the wall remains constant and does not vary with time. However, the temperatures at
different locations in the wall, though constant, will be different. Heat flows through the wall
(system) even after attainment of steady state, at a constant rate as different locations / points in
the wall have different temperatures. It is, therefore, obvious that in the steady state a system
exchanges mass, heat or work with its surroundings, even though the properties of the system do
not vary with time, owing to the presence/existence of a driving force of one kind or another.
(The temperature difference acts as a driving force in heat transfer, whereas the concentration
difference acts as a driving force in mass transfer.)
It is important not to confuse steady-state conditions with equilibrium conditions. In a steady
state, the properties at all points within the system are constant - they do not vary with time, but
they are different at different points. In a state of equilibrium also the properties do not vary /
change with time, but instead of different at different points, they are uniform within the system -
they are same at all points. In a steady state, the system can exchange mass, heat or work with the
surroundings since driving forces are not balanced while in a state of equilibrium, the system
cannot exchange mass, heat or work with the surroundings since in the equilibrium state, all
potentials / driving forces are balanced (driving forces are absent).
THERMODYNAMIC PROPERTIES OF A SYSTEM
Properties of a system are characteristics of the system which describe its physical condition,
e.g., volume, pressure, temperature, etc. They depend on the state of a system and not on how the
state is reached / attained. They are macroscopic in nature, the coordinates to describe the state
of a system and the state variables of the system.
The properties of a system may be classified in two groups as extensive properties and
intensive properties.
• The properties of a system which depend on the mass / amount (quantity) of the system
are called extensive properties.
• An extensive property of a system is one whose value depends on the mass of the
system. These properties are additive.
Volume, internal energy, enthalpy, entropy and heat capacity are extensive properties.
Since volume is an extensive property, it is additive. When two systems having the same
temperature are mixed, then for a combination of these two systems, temperature still remains the
same as original but the volume will be equal to the sum of the volumes of two systems.
Chemical Engineering Thermodynamics - I 1.11 Introduction and Basic Concepts

• The properties of a system which are independent of the mass / amount of the system
are called intensive properties.
• An intensive property of a system is the one whose value does not depend on the mass
of the system. These properties do not depend on the size of the system and are not
additive.
Temperature, pressure, specific volume, density, specific heat, molar volume, viscosity,
thermal conductivity, molar enthalpy and entropy, boiling point, melting point and freezing point
are intensive properties.
It should be noted that an extensive property may become an intensive property when the
extensive property is specified per unit mass basis. Thus, volume is an extensive property; but
specific volume, which is the volume per unit mass, is an intensive property.
The properties of a system depend upon the state of the system and not upon the path or
process by which the state is reached. They describe / denote the current / present state / status of
the system with regard to the property under consideration and do not give any idea regarding the
previous history / previous state of the system.
It should be noted that the value of a property is the same at all points within the system.
Thus, it is meaningless to speak of the pressure or the temperature of a system unless the pressure
or temperature has the same value throughout the system.
STATE OF A SYSTEM
The thermodynamic state or, in brief, the state of a system is the condition of the system at
any particular instant of time as described or measured by its macroscopic properties, such as
temperature, pressure, volume and composition.
Each unique condition of a system, as defined by its macroscopic properties, is called a state.
State denotes the values of properties of a system at a particular instant of time. Therefore, when
all the properties of a system have definite values, the system is said to exist in a definite state.
A change in any one of these properties causes the system to change to a new state. Thus, the
state of a system is fixed by its macroscopic properties. Stated differently, all properties are state
or point functions.
Pressure, volume, temperature and composition are the most important thermodynamic
properties (also called as thermodynamic variables / state variables). In case of a simple
homogeneous system (a homogeneous system consisting of a single substance), the composition
is fixed (it is always 100%) and hence the state of the system depends on the temperature,
pressure and volume. The value of any one of these properties / variables depends upon the value
of other two properties / variables. These three variables / properties of a simple homogeneous
system of a definite mass are related to one another by the relationship called an equation of
state. The equation of state for 1 mole of an ideal gas is PV = RT, where V is the molar volume
and R is the gas constant. The two variables / properties, generally specified, are temperature and
pressure (to describe the state of the system. They are also called thermodynamic coordinates).
When two of these properties, say, temperature and pressure are specified, all other physical
Chemical Engineering Thermodynamics - I 1.12 Introduction and Basic Concepts

properties of a simple homogeneous system, such as molar or specific volume, viscosity,


refractive index, thermal conductivity, etc. are thereby definitely fixed.
For example, oxygen gas at a temperature of 300 K and a pressure of 101.325 kPa has a
definite density or specific volume, a definite viscosity, a definite thermal conductivity, etc.
(i.e., it has a definite set of properties).
The variables (also properties), such as temperature and pressure used to define the state of
the system are independent variables. The third variable, volume is called a dependent variable.
Thus, the thermodynamic state of a simple homogeneous system is completely defined / fixed by
specifying / fixing any two of three variables (properties), e.g., temperature, pressure and
volume. However, for more complex system, the number of properties or variables that must be
arbitrarily specified / fixed to define the state of the system may be different from two (decided
by the phase rule).
STATE VARIABLES
The state of a system changes with change in any of the macroscopic properties, so these
properties are called state variables. Pressure, temperature and volume are the most important
state variables.
STATE FUNCTIONS
The properties of a system are known as state functions as they have definite / fixed values
for a particular state of the system and do not depend on the way / manner in which the state is
reached / attained.
Pressure, volume, temperature, internal energy, enthalpy and entropy are state functions
(point functions). The change in any state function accompanying a change in the state of a
system depends on the initial and final states of the system and not on the path by which the state
is brought about. Please note that the differential of a state function is an exact or a complete
differential and it represents an infinitesimal or differential change in the property.
Exact differentials can be integrated between the appropriate limits. The integration of such a
differential gives a finite change / difference between two values of the property. If Z is a state
function, the finite change / difference, ∆Z, as the system changes from state A to state B, is
given by
∆Z = ZA – ZB
P2 U2

Therefore, ⌠
⌡ dP = P2 – P1 = ∆P and ⌠
⌡ dU = U2 – U1 = ∆U
P1 U1

Consider a system at an initial state (P1, T1). It is taken to a final state (P2, T2) through / by
three different paths :
(i) A - in one step,
(ii) B - in two steps and
(iii) C - in several steps as shown in Fig. 1.2.
Chemical Engineering Thermodynamics - I 1.13 Introduction and Basic Concepts

S
P2 (P2, T2) - Final state

Path B
R - Initial state (P1, T1)
Path C
A S - Final state (P2, T2)
th
Pa
R
P1
Path B
Path C Path B Path A

T1 T2

Fig. 1.2 : Different paths from R to S


The change in temperature, ∆T = T2 – T1 or the change in pressure, ∆P = P2 – P1 of the
system remains the same in all the three cases and are independent of the paths by which the
change is brought out. Therefore, P and T are state functions. State functions can be identified by
points on a graph.
PATH FUNCTIONS
Thermodynamic functions / physical quantities which depend on the path followed by a
system in reaching the final state from the given initial state are called path functions e.g., heat
and work.
The values of heat and work accompanying a given change in state vary with the path from
the initial state to the final state, i.e., the values of these quantities depend on the manner in
which the change is brought about. Work and heat, as they depend on path, cannot be identified
by points on a graph, but are represented by areas. As work and heat appear only when changes
are caused in a system by a process, they are called properties of the process. They are not
properties of a system.
The work done by a system when it expands is given by the product of the external pressure
and the increase in the volume of the system. The work done for a given volume change depends
on the external pressure, the value of which may change according to circumstances. Therefore,
the value of work done will not be determined solely by the initial and final states of the system,
but also on the manner in which the change is carried out. For example, the work done in
changing the volume of a gas by one cubic meter (1 m3) would be 100, 10 or 1 J according to as
the external pressure is 100, 10 or 1 N/m2.
The differential of a state function is an exact differential or a complete differential whereas
the differential of a path function is an inexact differential or incomplete differential. State
functions can be identified by points on a graph whereas path functions are identified /
represented by areas. The differentials of heat and work are referred to as infinitesimal quantities
Chemical Engineering Thermodynamics - I 1.14 Introduction and Basic Concepts

of heat and work. The integration of these differentials results in a finite quantity and not a finite
change. Therefore,
2


⌡ dQ or ⌠
⌡ dQ = Q or 1Q2 or Q1-2 and ≠ Q2 – Q1
1
2

Similarly, ⌠
⌡ dW or ⌠
⌡ dW = W or 1W2 or W1-2 and ≠ W2 – W1
1

PROCESS
• Whenever a system undergoes a change of state, it is said to undergo a process.
• A process is the means by which the given change in state is achieved (or the means by
which a system changes from one state to another).
• A process is the means by which some sort of change or transformation, physical or
chemical, takes place in the system under consideration.
During a process, the system changes from an initial state to a final state, through a series of
intermediate states. This series of intermediate states through which a system passes during a
process to reach the given final state from the given initial state is called the path of the process.
Thus, a process undergone by a system will result in / cause a change in at least one of the
following :
Pressure, temperature, volume, phase present and chemical composition of the components
constituting the system.
A process may be a non-flow process or flow process. A non-flow process is one in which a
fixed mass within the defined boundary is undergoing a change of state. For example, a
substance which is being heated in a closed vessel undergoes a non-flow process. Closed systems
undergo non-flow processes. A flow process is one in which mass is continuously entering and
leaving through the boundary of an open system. In a steady flow process, a certain mass from
the surroundings is crossing the boundary of the system at entrance and an equal mass from the
system crossing the boundary of the system at the exit, so that the total mass of the system
remains constant.
TYPES OF PROCESSES
Isothermal Process :
• A process which is carried out at a constant temperature is called as an isothermal
process.
• It is a process in which the temperature of the system remains constant.
• An isothermal process is defined as one which is carried out at a constant temperature,
at all the stages.
For an isothermal process : Change of temperature = 0.
Chemical Engineering Thermodynamics - I 1.15 Introduction and Basic Concepts

Adiabatic Process :
• A process which is carried out in such a way that no heat enters or leaves the system is
called as an adiabatic process.
• It is a process in which no heat leaves or enters the system, at any stage.
• An adiabatic process is defined as one in which no heat leaves or enters the system,
at any stage.
For an adiabatic process : Heat change = 0.
Isobaric Process :
• A process which is carried out at a constant pressure is called as an isobaric process.
• It is a process in which the pressure of the system is kept constant. For an isobaric
process, change of pressure = 0. For example, a reaction carried out in an open vessel.
Isochoric Process :
• A process which is carried out at a constant volume is called an isochoric process.
• An isochoric process is defined as one which is carried out at a constant volume.
• It is a process in which (during which) the volume of the system is kept constant.
For example, a gas phase reaction carried out in a sealed / closed vessel.
Cyclic Process :
• A cyclic process is said to have undergone by a system when after changing from some
initial state to some other state, the system is returned back exactly to its original/initial
state.
• A process, or a series of processes, undergone by a system as a result of which the
system is returned / restored exactly to its initial / original state is called as a cyclic
process or cycle.
Suppose that pure alcohol and pure water are mixed to prepare a 50% alcohol solution by
pouring the alcohol into the water. If this 50% alcohol solution is separated by distillation or by
any other suitable technique into its original pure components, each in its original / initial state,
then the system originally / initially consisting of this pure alcohol and pure water has completed
a cyclic process. It should be noted that the system has returned to its initial state as a result of
such a cyclic process, while the surroundings are not necessarily returned to their original state /
condition.
ENERGY
Since thermodynamics deals with energy and its transformation, the concept of energy is
essential in the understanding of thermodynamics. Energy is the capacity / ability of a body to do
work. The forms of energy can be classified as :
Chemical Engineering Thermodynamics - I 1.16 Introduction and Basic Concepts

(i) Forms of energy related to the system : These include energy possessed by material
of the system - kinetic energy, potential energy, pressure / flow energy, internal energy, surface
energy and magnetic energy.
(ii) Forms of energy associated with the process : These include energy produced or
transferred by the processing - heat and work.
Kinetic energy : The energy possessed by a body by virtue of its motion is called the kinetic
energy. If a body of mass m is moving at a velocity u, then the kinetic energy of the body is
1
K.E. = 2 mu2 ... (1.1)

Potential energy : The energy possessed by a body by virtue of its position in relation to
some datum plane is called the potential energy. If a body of mass m is at an elevation Z above
the ground, then the potential energy of the body is given by
P.E. = mgz ... (1.2)
The work done on a body of mass m in raising its elevation from z1 to z2 is given by
W = F (z2 – z1) = mg (z2 – z1) = mg ∆z
The work done on a body in raising its elevation is stored in the body as energy and this
energy can be recovered from the body by lowering its elevation.
Since m, u and z are macroscopic quantities, both the kinetic and potential energies are called
macroscopic modes of energy. The sum of the kinetic energy and potential energy of a body is
called the mechanical energy of a body.
Flow / Pressure energy : The flow or pressure energy (PV) is the product of the pressure
and volume and is associated with flowing streams under pressure.
Internal energy : The internal energy of a substance is the energy of the molecules
constituting the substance and not the energy possessed by the substance due to / by virtue of its
macroscopic movement or position.
It is the energy a substance possesses because of the motion and configuration of its
molecules.
The energy stored in a system because of the motion and configuration of the molecules
constituting the system is called its internal energy.
This is different from the kinetic energy and potential energy a substance possesses because
of the macroscopic motion and position of the substance as a whole which are thought of as
external forms of energy.
The molecules constituting a substance possess kinetic energy of rotation, translation and
internal vibration. In addition to the kinetic energy, the molecules of a substance possess
potential energy due to the forces of attraction among them. The sum of these molecular kinetic
and potential energies (microscopic modes of energy) is called the internal energy of the system.
It is designated by the symbol U.
Chemical Engineering Thermodynamics - I 1.17 Introduction and Basic Concepts

The addition of heat to a substance increases the internal energy of the substance as added
heat increases its molecular activity.
The internal energy of a system is a function of the state of the system, i.e., it is a point / state
function and hence a property of a system. It is not possible to determine its absolute value but,
fortunately, in thermodynamic analysis the absolute value is of no significance and what is
needed is the change in internal energy accompanying a given process and this is a measurable
quantity.
The total energy E of a system is equal to the internal energy (U), the energy due to the
position of the system (P.E.) and the energy due to the motion of the system as a whole (K.E.).
Two forms of energy - surface energy and magnetic energy are usually negligible in magnitude
compared with the other forms and thus are not included in E.
Therefore, E = U + K.E. + P.E. ... (1.3)
An insulated spherical vessel containing a certain gas has a certain internal energy. It can
have a potential energy equal to mgz due to its position and if the whole body is moving with a
1
certain velocity u, it also has a kinetic energy equal to 2 mu2.

E given by Equation (1.3) is the energy stored in a system, while heat and work are the forms
of energy in transit, which are observed at the boundaries of a system. The magnitudes of heat
and work depend on the path followed by the system during a change of state.
Processes are often encountered in which there are no changes in the kinetic and potential
energies of the system and in such cases ∆E = ∆U.
∆E = ∆U + ∆ KE + ∆ PE
For ∆ KE and ∆ PE ≈ 0,
∆E = ∆U
THERMODYNAMIC EQUILIBRIUM
A system is said to be in a state of thermodynamic equilibrium if the properties on a
macroscopic scale, i.e., the macroscopic properties (such as temperature, pressure, density,
composition or concentration of species) are uniform throughout the system and do not undergo
any change with time.
The state in which the macroscopic properties of a system are uniform throughout the system
and do not undergo any noticeable change with time, under a given set of conditions, is said to be
a state of equilibrium. This implies a balance of all potentials that tend to cause a change.
At equilibrium, all potentials tending to cause a change are balanced and hence there is no
further tendency for a change to occur (all tendency towards change ceases). Thus, equilibrium
refers to the absence of any tendency for a change to occur. Changes in a system are always
caused by the action of some form of driving force, or potential difference. The driving force or
potential difference in heat flow / transfer is temperature difference (temperature is one of the
Chemical Engineering Thermodynamics - I 1.18 Introduction and Basic Concepts

potentials). The driving force (or potential difference) in fluid flow is difference of pressure or
head and in mass transfer it is concentration difference (true driving force - chemical potential).
Thus, equilibrium implies the absence of driving force between the two interacting bodies (as at
equilibrium any tendency towards change ceases).
The thermodynamic equilibrium of a system implies that state wherein the system is in
mechanical, thermal and chemical equilibrium. In a state of thermodynamic equilibrium, the
system does not exchange heat as well as work with its surroundings (physical change) and does
not undergo any chemical change. However, at the microscopic level, the conditions are not
static even after the attainment of equilibrium.
A system is said to be in thermal equilibrium / in a state of thermal equilibrium if there is
no flow of heat from one portion of the system to another. This is possible only if the
temperature remains the same / uniform throughout the system. Thus, thermal equilibrium
denotes uniformity of temperature or the absence of temperature gradient. In a state of thermal
equilibrium, there is no temperature gradient within the system and there is no difference in
temperature (driving force) between the system and its surroundings.
A system is said to be in a state of mechanical equilibrium if there is no imbalance of
forces within the system. In a state of mechanical equilibrium, the system has a uniform pressure
and there is no pressure difference between two interacting systems. Mechanical equilibrium
denotes uniformity of pressure or the absence of unbalanced of forces.
A system is said to be in a state of chemical equilibrium if the composition of the
components present in the system is uniform throughout the system and there is no tendency for
the chemical reaction to occur within the system. Chemical equilibrium denotes equality of
chemical potential.
The criteria of equilibrium are :
Equality of temperature : Thermal equilibrium
Equality of pressure : Mechanical equilibrium
Equality of chemical potential : Chemical equilibrium
If the system is simultaneously in a state of thermal equilibrium, mechanical equilibrium and
chemical equilibrium, then it is said to be in a state of thermodynamic equilibrium.
A system in a state of thermodynamic equilibrium satisfies the conditions / criteria of
thermal, mechanical and chemical equilibrium. A system in thermal equilibrium has uniform
temperature, so a unique / single value can be assigned to the temperature of the system. Hence,
single values can be assigned to the properties of a system only when it is in a state of
thermodynamic equilibrium.
Thermodynamics deals with systems which are in thermodynamic equilibrium or idealized to
be in thermodynamic equilibrium.
REVERSIBLE AND IRREVERSIBLE PROCESSES
The processes we normally come across can be divided into two classes :
(i) Reversible processes and
(ii) Irreversible processes.
Chemical Engineering Thermodynamics - I 1.19 Introduction and Basic Concepts

A reversible process is one


(i) in which the driving force causing the process is always infinitesimally greater than the
opposing force (i.e., the force resisting the process),
(ii) which is carried out extremely / infinitesimally slowly,
(iii) which is free of dissipative effect (friction, viscosity, inelasticity and electrical
resistance) of any form and
(iv) which can be stopped and reversed at any stage so that both the system and its
surroundings are restored to their respective initial / original states.
A process which is carried out with a finite driving force is called an irreversible process.
Since a reversible process is carried out infinitely slowly, with an infinitesimal driving force,
the system is at all times infinitesimally near a state of thermodynamic equilibrium and thus
every state passed through by the system is an equilibrium state.
A reversible process is thus a succession or series of equilibrium states. As these states can
be described by thermodynamic properties (e.g., P, V, etc.), any reversible process can be
represented on a thermodynamic diagram / can be plotted on thermodynamic coordinates.
If a process occurs in such a way that the system passes through non-equilibrium states (non-
equilibrium intermediate states) though the initial and final states of the system are equilibrium
states, it cannot be plotted on thermodynamic coordinates/cannot be represented on a
thermodynamic diagram since because intermediate non-equilibrium states cannot be described
by thermodynamic properties. Such a process is called an irreversible process.
When the system is in equilibrium at all the intermediate states between the initial and final
states (which are also equilibrium states) only then the path followed by the system can be
defined / can be traced. The path of a reversible process between the initial and final states is
always represented by a continuous line as the path followed is certain, whereas the path of any
irreversible process between the two states is always represented by a dotted line as the path is
uncertain (the path of such a process cannot be traced).

(a) (b)
Fig. 1.3 : Reversible and irreversible processes
Chemical Engineering Thermodynamics - I 1.20 Introduction and Basic Concepts

The reversible and irreversible processes are represented on a P-V diagram as shown in
Fig. 1.3. The dotted line has got no meaning otherwise because it can be drawn in any
way/fashion.
A reversible process is an ideal or imaginary one since in this process all sources of energy
dissipation are eliminated. Since reversible processes take place infinitesimally slowly, they
would require infinite time for their completion. Therefore, reversible processes are not
practicable. Inspite of impracticability of reversible processes, the study of these processes is of
great value / importance since we get idea regarding the maximum efficiency obtainable in any
given change. Thus, the ideal to be aimed at is known (the limit to the performance of an actual /
real system).
Irreversible processes are those in which the condition of thermodynamic equilibrium is not
satisfied and dissipative effects (e.g., fluid friction) are present. Since these processes take place
fast with a measurable speed, finite time would be required for their completion. All actual
processes that occur in nature are irreversible since they take place with finite driving forces
(e.g., ∆T for heat transfer) between parts of the system or between the system and its
surroundings. The irreversibility of a process may be caused due to lack of equilibrium during
the process and / or presence of dissipative effects (friction, viscosity, electrical resistance).
Examples of reversible process : motion without friction, heat transfer with no temperature
difference between the system and its surroundings.
Examples of irreversible process : motion with friction, unrestricted / free expansion
(of a gas), rusting of iron, mixing, heat transfer with a finite temperature difference.
The isothermal vaporization carried out in the following manner provides a simple example
of a reversible process. Imagine a cylinder, containing a liquid in equilibrium with its vapour,
placed in a constant temperature bath, i.e., a large thermostat. Further consider that the cylinder
is closed by a frictionless piston. If we increase the external pressure acting on the piston by an
infinitesimally small amount, some of the vapours will condense. Since the condensation will
occur extremely slowly, the heat liberated (the latent heat of condensation) will be taken up by
the thermostat. Therefore, the temperature of the system will not rise and the pressure above the
liquid will remain constant. Although the process of condensation is taking place, the system is
always (i.e., at every instant) infinitesimally near a state of thermodynamic equilibrium.
Similarly, if we make the external pressure smaller than the vapour pressure of the liquid by an
infinitesimally small amount, the liquid will vaporize extremely slowly. In this case also
temperature and pressure will remain constant. Since vaporization is taking place, the system is
changing, but the process may be regarded as a succession of thermodynamic equilibrium states.
If we carry out the vaporization or condensation rapidly, by suddenly decreasing or increasing
the external pressure, these actions will lead to temperature and pressure gradients within the
system and thermodynamic equilibrium will be disturbed. The phase change operations will not
involve a continuous succession of equilibrium states and hence, such operations are not
reversible.
Chemical Engineering Thermodynamics - I 1.21 Introduction and Basic Concepts

REVERSIBLE WORK OF EXPANSION


Work done by a force on a body is defined as the product of the force and the displacement
of the body in the direction of the applied force. Mathematically, this can be written as
dW = F dL ... (1.4)
where dW is the differential amount of work done for dL, the differential displacement in the
direction of the applied force, F.
However, in thermodynamics an important type of work is one which is associated with a
change in the volume of a fluid - work due to a change in the volume of a fluid against an
external pressure (mechanical work/pressure-volume work/expansion or compression
work/displacement work).
Mechanical work is performed when a system changes its volume against an opposing
pressure. Consider that a gas in a piston - cylinder assembly expands against a constant external
pressure, Pex and its volume changes by dV as the piston moves through a distance dL. Here
pressure of the gas is greater than the pressure on it of the surroundings, Pex.
The external force which restrains expansion of the gas acting on the piston is given by
F = Pex A ... (1.5)
where A is the area of the piston normal to the force.
This force has moved through the distance (– dL). Therefore, the differential amount of work
done by the applied force is given by
dW = Pex · A × (– dL) = – Pex A dL
dW = – Pex dV ... by the force ... (1.6)
where dV = A dL = Differential change in volume of the system
The work done by the system (i.e., the gas in the piston - cylinder assembly) on the
surroundings (i.e., applied force) is equal to the negative work done by the applied force on the
system.
Therefore, the work done by the system is given by
dW = – [– Pex dV] = Pex dV ... (1.7)
Differential change
∴ Differential = External pressure ×  in volume 
 work done 
 of the system 
If the volume of the gas changes from the initial volume V1 to the final volume V2, then
Equation (1.7) can be integrated.
V2


⌡ dW = ⌠
⌡ Pex dV
V1
V2

W = ⌠
⌡ Pex dV ... (1.8)
V1
Chemical Engineering Thermodynamics - I 1.22 Introduction and Basic Concepts

A general expression for the work of expansion accompanying a reversible process may be
readily derived.
Consider a gas enclosed in a cylinder fitted with a weightless and frictionless piston,
as shown in Fig. 1.4 (a) undergoing a reversible expansion process. Let P be the pressure of the
gas (the system). For a reversible expansion of the gas, the external pressure must be
infinitesimally smaller than the pressure of the gas. Therefore, the external pressure is
P – dP, where dP is a very small quantity. As the external pressure P – dP is infinitesimally
smaller than the gas pressure (pressure of the system), the gas will expand by an infinitesimal
volume dV (the volume of the gas changes from V to V + dV). The work done by the gas when
its volume increases by an infinitesimal amount dV is equal to the external pressure times the
volume change [as given by Equation (1.7)].
dW = External pressure × Change in volume
= (P – dP) dV
= P dV – dP dV
Neglecting the product dP dV, as both the quantities are very small / infinitesimal (and
hence, the product dPdV), the above equation becomes
dW = P dV ... (1.9)
Equation (1.9) gives the differential amount of work done by the gas.

1
P1

P Equilibrium state
P
Gas
P
External pressure
P2 2
dV

V1 V V2

(a) (b)
Fig. 1.4 : (a) Work done by a system during reversible expansion
(b) P-V diagram for a reversible expansion of a gas
The total work done by the gas (the system) in the expansion process as the piston moves
from position 1 to position 2 during which the volume is changing from V1 to V2 is given by
2 V2

W = W1-2 = ⌠
⌡ P dV = ⌠
⌡ P dV ... (1.10)
1 V1
Chemical Engineering Thermodynamics - I 1.23 Introduction and Basic Concepts

Since the external pressure is infinitesimally smaller than the pressure of a gas, the piston
moves infinitesimally slowly and thus every state passed through by the system from V1 to V2 is
an equilibrium state. Every equilibrium state is characterized by uniformity of thermodynamic
properties throughout the system and the system is having a known / well defined set of
thermodynamic properties. Therefore, the path followed by the system is well defined and can be
drawn on a P-V diagram and the area under the curve gives the value of the integral P dV. This is
true only for a reversible process because the intermediate states passed through by the system
during this process are equilibrium states (since the acting force is infinitesimally different from
the opposing force and the system thus passes through equilibrium states). Therefore, for any
reversible process, the work done is evaluated using the relation W = ⌠
⌡ P dV from the properties
of the system. On the other hand, in case of an irreversible process, the acting force is far
different from the opposing force and the system passes through non-equilibrium states which are
indeterminate, i.e., not exactly known or defined. Therefore, for any irreversible process, the
work done, W cannot be evaluated using the relation W = ⌠
⌡ P dV. Thus the integration
⌠ P dV is performed only on a reversible path.

The work done by the gas (the system) on the surroundings is given by the area under the
curve (the path 1 - 2) on a P-V diagram as shown in Fig. 1.4 (b).
The work done is the area enclosed by the curve and the two ordinates at the initial and final
volumes (V1 and V2).

P III
II
I
2

V1 V2
V
Fig. 1.5 : Work - dependent on the path followed
It is possible to take a system from state 1 to state 2 along various paths, such as I, II or III
(Refer Fig. 1.5). The work done along each path is the area under each curve. It can be observed
that the area under the path-I is different from that under the path-II which is still different than
the area under the path-III (curve-III). The work of expansion will thus clearly be a variable
quantity.
Chemical Engineering Thermodynamics - I 1.24 Introduction and Basic Concepts

Thus, the amount of work done in each case is not a function of the end states (1 and 2) of
the process but depends on the path the system (the gas) follows in going from state 1 to state 2.
Because of this, work (work interaction) is a path function and not a point / state function and its
differential – dW is not exact. – dW is an inexact differential.

Note that Equation (1.10) is applicable to both expansion and compression. When there is an
expansion, V2 is greater than V1 and hence, the work done by the system (W) is positive. On the
other hand, when there is a compression, V2 is less than V1, and W obtained by Equation (1.10)
is negative. This is expected, as in compression, work is done on the system.

When the work is done by a system on the surroundings, by the convention, it is taken as
positive, and when the work is done on the system by its surroundings, it is taken as negative.
Work output of a system = + W and work input to the system = – W. When a system does work,
it loses energy and when the work is done on the system, it gains energy.
Surroundings Surroundings
W

Boundary
System System System gains
energy as
System loses work
W
energy as work
(a) W is +ve (b) W is -ve

Surroundings Surroundings
Q
System gains
energy as
System heat System
System loses
Q energy as
heat
(c) Q is +ve (d) Q is -ve

Fig. 1.6 : Work and heat interactions between a system and its surroundings
In integrating Equation (1.10), it is important to use consistent units. For example, if V is
expressed in terms of cubic meters (m3), then P must be in N/m2 (Pa), so that the product of
pressure and volume will have the dimensions of work, namely newton.meters in the SI system.
The unit of work is N.m or J.
In some situations, W ≠ 0, even though dV = 0. For example, a liquid contained in a rigid
vessel, which is stirred by a rotating paddle wheel. In this case, since the walls of the vessel are
rigid, dV = 0, but the work is done on the system by the paddle wheel.
Chemical Engineering Thermodynamics - I 1.25 Introduction and Basic Concepts

In some situations, W = 0 even though dV ≠ 0. For example, in the process of free expansion
of a gas, the expansion of a gas against vacuum is called free expansion. Consider a vessel
divided into two compartments by a partition. One compartment contains a gas at a given
pressure, while the second one is completely evacuated - a gas separated from the vacuum by a
partition (Fig. 1.7). If the partition is removed, the gas expands and fills the entire vessel. In this
case, as the expansion of the gas is not restrained by an opposing force, it is called free
expansion. In the process of free expansion, as the gas expands into a vacuum, that is, against no
external pressure (no restraining force has moved through any distance), the work of expansion is
zero (W = 0).
Partition

Gas Vacuum
P

Fig. 1.7 : Free expansion of a gas


HEAT
Heat is a form of energy which is transferred across a boundary due to the existence of a
temperature difference.
Heat is the energy in transit (like work) between a hot source and a cold receiver. It appears
at the boundary of a system, while a change is taking place within the system. The amount of
heat transferred in a process depends on the way in which the process is carried out (it depends
upon the intermediate states through which the system passes, i.e., path). Therefore, just as work,
heat is a path function and not a state function.
The symbol Q is used for the amount of heat transferred. The unit of heat in the SI system is
joule (J).
Heat flow into a system from the surroundings, by convention, is regarded as positive. Thus,
Q is positive when the system gains energy as heat. Heat flow out of a system, that is, heat given
up by the system to its surroundings, i.e., heat transferred from the system to its surroundings is
taken as negative (since there is a corresponding decrease of energy of the system).
Heat added to a system = + Q ... taken as a positive quantity
Heat given up/rejected by a system = – Q ... taken as a negative quantity
A system exchanges energy with its surroundings in the form of heat and work, i.e., a system
interacts with its surroundings by transfer of energy as heat and work. These energy interactions
bring about changes in the properties of the system. Both heat and work are energy in transit and
can be identified only when the process is in progress (only when a certain change is taking
Chemical Engineering Thermodynamics - I 1.26 Introduction and Basic Concepts

place). Heat and work are path functions and their differentials dQ and dW are inexact
differentials. Both are boundary phenomena and are associated with a process, not a state and
have no meaning at a state. They are not properties of a system but are properties of a process.
According to the convention, Q and W are the heat absorbed and work done by the system.
Therefore, when
Q is + ve : heat has been added to the system
Q is – ve : heat has been removed from the system
(heat has been given up / rejected by the system)
W is + ve : Work has been done by the system
W is – ve : Work has been done on the system
Work done by the system = Work output of the system = Work produced
Work done on the system = Work required = Work input
If Q = 500 kJ ... means heat added / absorbed by the system is 500 kJ
If Q = – 400 kJ ... means heat removed from the system is 400 J
If W = 300 kJ ... means work output of the system is 300 kJ
If W = – 450 kJ ... means work required is 450 kJ
TEMPERATURE AND ZEROTH LAW OF THERMODYNAMICS
Temperature is a thermal state of body, which distinguishes a hot body from a cold body.
Temperature is a measure of the degree of hotness or coldness of a body. A body which is
hot is said to have a higher temperature than one which is cold. If a hot body is placed in contact
with a cold body, then after a long time in contact, they attain a common temperature (both will
be at the same temperature) and the two bodies are said to attained a state of thermal
(or temperature) equilibrium or said to exist in thermal equilibrium. The temperature of the hot
body has decreased, while that of the cold body has been raised (because of energy transfer as
heat from the hotter body to the colder body) until at equilibrium the two bodies have
equal/identical temperature.
In order to do a precise quantitative measurement of temperature, an instrument or device
known as a thermometer is used. For the measurement of temperature by a thermometer, a certain
property of the thermometer which changes with temperature (the degree of hotness) is selected
and the changes in this property may be taken as an indication of the changes in temperature.
Such a property is usually called thermometric property. This property in a mercury thermometer
is the length of the mercury column in a capillary tube (the variation of volume with temperature,
which is observed by the change in length of the mercury column), in a thermocouple it is the
Chemical Engineering Thermodynamics - I 1.27 Introduction and Basic Concepts

e.m.f. generated at the junction of two dissimilar metal wires and in case of a resistance
thermometer it is the electrical resistance of a metal wire.
When a body A is brought in thermal contact with a body B, heat flows from one to the other
(heat flow is caused by the action of a potential or driving force called temperature) and
eventually thermal equilibrium is established between them (between A and B). When two
bodies are in thermal equilibrium, the temperatures of both the bodies are equal.
If a body A is in thermal equilibrium with a body C and a body B is in thermal equilibrium
with the body C, then the two bodies A and B are also in thermal equilibrium with each other.
If two bodies are each in thermal equilibrium with a third body, then they (the two bodies)
are in thermal equilibrium with each other.
This is known as the zeroth law of thermodynamics. This law was formulated by
R.H. Fowler in 1931. However, at that time, the first and second laws were already existed.
Due to this, it was named as zeroth law of thermodynamics so that it precedes the first and
second laws to form a logical sequence. The zeroth law of thermodynamics forms the basis for
the measurement of temperature.
Suppose the body C in the previous illustration is a mercury thermometer. When this
thermometer is placed in contact with the body A until thermal equilibrium is attained, it has a
certain height / length for the mercury column in the capillary (which depends on the temperature
of the body A). If this thermometer is brought in contact with the body B and indicates the same
height for the mercury column (when thermal equilibrium is attained between them), then we say
that the bodies A and B are at the same temperature, without actually bringing the bodies A and
B in thermal contact.
To devise a temperature scale, that is, to assign a numerical value to each level of
temperature, it is necessary to have reference states with known temperatures. The reference
states chosen are the ice point and the steam point (the normal boiling point of water - the boiling
point of water at standard atmospheric pressure). In the Celcius scale of temperature (on the
Celcius temperature scale), the ice point (the freezing point of water / the melting point of pure
ice at standard atmospheric pressure) has been assigned a value of 0°C and the steam point has
been assigned a value of 100°C. The interval between these two fixed points is divided into
hundred equal parts and each part is designated as one degree Celsius. Since the thermometric
properties of different thermometers do not vary in an identical manner with temperature,
thermometers, all of which have been standardized at 0°C and 100°C, may indicate different
temperatures when they are in thermal equilibrium with the same body at an intermediate point.
In order to overcome this difficulty, it is necessary to establish a temperature scale that does not
depend on the nature of the thermometric fluid.
IDEAL GAS TEMPERATURE SCALE
An ideal gas is a gas which obeys the relationship PV = RT, where R is the universal gas
constant. At sufficiently low pressures (as P → 0), all real gases behave like an ideal gas. When
Chemical Engineering Thermodynamics - I 1.28 Introduction and Basic Concepts

gases approximate ideal behaviour, i.e., at very low pressures, the differences in their
thermometric properties disappear. This fact enables us to devise a temperature scale which is
independent of the nature of the thermometric fluid (the gas). The ideal gas temperature scale is
based on the fact that the pressure exerted by a fixed / definite quantity of gas is directly
proportional to the temperature at constant volume.
Consider a constant volume gas thermometer as shown in Fig. 1.8. Suppose a certain quantity
of gas is enclosed in the constant volume bulb B. This bulb is connected to one limb of the
mercury manometer M via the capillary tube C. The other limb of the mercury manometer is
open to the atmosphere. The volume of the gas is calibrated upto the reference mark R.
By raising or lowering the limb of the manometer which is open to the atmosphere (adjustable
limb), the volume of the gas in the bulb B can be maintained constant at the reference mark R.

C
h Adjustable limb/arm

R N

Mercury
M
B
Gas

Flexible tube

Fig. 1.8 : Constant volume gas thermometer


The pressure in the bulb is used as a thermometric property. It is given by the relation
P = P0 + hρg
where P0 is the atmospheric pressure and ρ is the density of mercury.

When the gas bulb is immersed in a bath whose temperature is to be measured, the bulb in
some time comes in thermal equilibrium with the bath. As the gas is heated, it expands and
pushes the mercury downward. The adjustable limb is then so adjusted that the mercury again
touches the mark R, i.e., the gas is filled upto the mark R only and the difference in mercury level
h is measured, with this h, the pressure of the gas in the bulb is estimated.
Suppose that the thermometer is now immersed in a bath at the triple point (the temperature
at which the solid, liquid and vapour phases of water coexist in equilibrium). The triple point
temperature Tt is assigned a value of 273.16 K, where K is the abbreviation for degree kelvin
(the SI unit of temperature).
Chemical Engineering Thermodynamics - I 1.29 Introduction and Basic Concepts

The temperature T of the bath corresponding to the pressure P of the gas (estimated from the
measured h) is given by
T P
Tt = Pt

T = 273.16 P 
P
or ... (1.11)
 t
Doing the measurements at P and Pt, the temperature T of the gas can be obtained using
Equation (1.11). If we do the measurements using different quantities of gas in the bulb, calculate
the temperatures using Equation (1.11) and plot these temperatures against Pt, then the
dependence of T on Pt for different gases is as shown in Fig. 1.9. [Surround the bulb, containing
a certain quantity of a gas, by water at its triple point and measure Pt. Then surround the bulb
with, say, condensing steam at 1 atm and measure the gas pressure P and calculate the
corresponding temperature. By successively reducing the quantity of gas repeat the above
procedure for different quantities and construct a plot of T v/s Pt. Repeat this procedure for
different gases.] If the lines, obtained by plotting T v/s Pt, are extrapolated to the y-axis where
Pt = 0, all gases indicate the same temperature (although the readings of a constant volume gas
thermometer depend upon the nature of the gas). This is so, because only when Pt tends to zero,
all gases behave like an ideal gas.

O2

Air

373.15

N2

T H2
(in K)

O
Pt, mm Hg

Fig. 1.9 : Plot of T v/s Pt for a constant volume gas thermometer (for steam point)
INTERNATIONAL PRACTICAL TEMPERATURE SCALE (IPTS - 1968)
In actual practice it is frequently necessary to measure temperature much higher than the
normal boiling of water and lower than the normal freezing point of water. In order to calibrate
thermometers for such temperatures, an international practical temperature scale has been
established. This scale is based on reproducible equilibrium temperatures or fixed points to
which numerical values have been assigned. The fixed points of the International Practical
Temperature Scale adopted in 1968 (abbreviated as IPTS - 68) are given in Table 1.1.
Chemical Engineering Thermodynamics - I 1.30 Introduction and Basic Concepts

Table 1.1 : Fixed points of IPTS - 68


Equilibrium state Assigned value of temperature in °C
Triple point of hydrogen – 259.34
Normal boiling point of hydrogen – 252.87
Triple point of oxygen – 182.97
Normal boiling point of oxygen – 218.789
Triple point of water 0.01
Normal boiling point of water 100.00
Normal boiling point of sulphur 444.60
Normal melting point of antimony 630.50
Normal melting point of silver 960.80
Normal melting point of gold 1063.00
• The triple point represents an equilibrium state between solid, liquid and vapour phases
of a substance - the state at which all the three phases of a substance coexist in
equilibrium with one another.
• Normal boiling point is the temperature at which the substance boils at standard
atmospheric pressure of 760 mm Hg (1 atm).
• Normal melting point is the temperature at which the substance melts at standard
atmospheric pressure of 760 mm Hg.
PRELIMINARY MATHEMATICAL BACKGROUND
[Required in deriving thermodynamic property relations ⇒ Chapter 5].
Exact Differential :
If a variable quantity z has a fixed value for every pair of values of x and y, then z is called a
function of two independent variables x and y.
If z is a function of x and y, this may be expressed as
z = f (x, y)
e.g.‚ P = f (V‚ T) ⇒ P = RT – a 
 (V – b) V2
The total differential of z can be written in terms of partial derivatives of z with respect to
x and y as given below.
∂z ∂z
dz =   dx +   dy ... (1.12)
∂xy ∂yx
 ∂  implies partial differentiation with respect to y keeping x constant.
∂y
 x
Chemical Engineering Thermodynamics - I 1.31 Introduction and Basic Concepts

∂z implies partial derivative / differentiation of z with respect to x keeping y


∴ ∂x
 y
constant. Partial derivatives mean that all other variables except the one that with respect to
which the differential coefficient is sought are held constant.
A differential equation of the type given by Equation (1.12) is known as an exact or perfect
differential equation. The differential dz or the differential of z is known as an exact differential.
∂z ∂z
If   = M and   = N, then Equation (1.12) becomes
∂xy ∂yx
dz = M dx + N dy ... (1.13)
where M and N are functions of x and y.
Differentiating M partially with respect to y (i.e., y differentiating M with respect to y
keeping x constant) and N with respect to x keeping y constant gives
∂M =  ∂ ∂z  denoted by ⇒ ∂2z
 ∂y   
 x ∂y ∂xyx ∂y ∂x

∂2z
= ... z is first differentiated with respect to x and then
∂y ∂x
with respect to y
∂N =  ∂ ∂z  = ∂2z
 ∂x   
 y ∂x ∂yxy ∂x ∂y

 ∂2z = ∂2z ,
Since the order of differentiation makes no difference in the final result  
∂y ∂x ∂x ∂y
we have
∂M = ∂N
 ∂y    ... (1.14)
 x  ∂x y
This is the condition of exact / perfect differential. dz is an exact differential only if
Equation (1.14) is satisfied for any function z. If the condition given by Equation (1.14) is
satisfied, then dz is said to be an exact differential and z is called a point function or state
function.
Equation (1.14) is called the Euler reciprocal relation. It applies only to state / point
functions, such as P, V, U, H, etc.
The differential of a state function is an infinitesimal change in the property. Exact
differentials can be integrated between the appropriate limits and results in a finite change in the
Chemical Engineering Thermodynamics - I 1.32 Introduction and Basic Concepts

U2

value of the property ⇒ ⌠


⌡ dU = U2 – U1 = ∆U. The differentials of heat and work as they are
U1

not state properties are inexact differentials and are referred to as infinitesimal quantities and not
changes. Integration of inexact differentials yield a finite quantity and not a finite change
2


⌡ dW = W1-2 ≠ W2 – W1.
1

The value of ⌠
⌡ dz between the two end conditions depends only on the values of z at these
 2


conditions i.e.‚ ⌡ dz = z2 – z1 and not on the path followed to bring about the change from
 1 
the initial condition to the final condition.
(1) Verify whether dz = (x2 – y2) dx + (y2 – 2xy) dy is an exact differential or not.
dz = (x2 – y2) dx + (y2 – 2xy) dy
Here M = x2 – y2 and N = y2 – 2xy
∂M = 0 – 2y = – 2y and ∂N = 0 – 2y = – 2y
Therefore,  ∂y   ∂x 
 x  y
∂M = ∂N
∴  ∂y   
 x  ∂x y
Hence, dz is an exact differential.
(2) Verify whether dz = (x2 y + 2y3) dx – (2x3 + 2xy2) dy is an exact differential or not.
dz = (x2 y + 2y3) dx – (2x3 + 2xy2) dy
Here M = x2 y + 2y3 and N = – (2x3 + 2xy2)
∂M = x2 + 6y2 and ∂N = – 6x2 – 2y2
 ∂y   ∂x 
 x  y
∂M ≠ ∂N , dz is not an exact differential.
As  ∂y   
 x  ∂x y
(3) For an ideal gas, show that dV is an exact differential.
V = f (P, T)
The total differential of V is
∂V ∂V
dV =   dP +   dT ... (1.15)
 ∂P T  ∂T P
RT
For an ideal gas, PV = RT so that V = P
Chemical Engineering Thermodynamics - I 1.33 Introduction and Basic Concepts

Differentiating V w.r.t. P at constant T as well as w.r.t. T at constant P gives


∂V = –RT and ∂V = R
 ∂P   ∂T 
 T P2  P P
Substituting these values in Equation (1.15), we get

dV = – P2  dP +  P  dT
RT R
   
–RT R
M = P2 and N = P

Differentiating M partially with respect to T and N with respect to P, we get


∂M –R ∂N –R
 ∂T  = P2 and  ∂P  = P2
 P  T
∂M = ∂N , = – R, dV is an exact differential.
As  ∂T   
 P  ∂P T  P2 
(4) For a van der Waals gas, show that dP is an exact differential.
If P is a function of V and T, it is expressed mathematically as
P = f (V, T)
The total differential of P is
 ∂P  ∂P
dP =   dV +   dT ... (1.16)
 T
∂V ∂TV
The van der Waals equation of state is
P + a  (V – b) = RT
 V 2
Rearranging gives
RT a
P = V – b – V2
Differentiating P w.r.t. V at constant T as well as P w.r.t. T at constant V, we get
 ∂P  –RT 2a
∂V = (V – b)2 + V3
 T
∂P = R
and ∂T
 V (V – b)
∂P  ∂P 
Substituting the values of   and   obtained, Equation (1.16) becomes
 V
∂T ∂VT
dP = (V – b)2 + V3 dV + V – b dT
–RT 2a R
   
–RT 2a R
∴ M = (V – b)2 + V3 and N = V – b
Chemical Engineering Thermodynamics - I 1.34 Introduction and Basic Concepts

Differentiating M partially with respect to T and N with respect to V, we get,


∂M =  ∂  –RT + 2a  = –R + 0 = –R
 ∂T  ∂T (V – b)2 V3
 V  V (V – b)2 (V – b)2

∂N =  ∂  R  = –R
∂V  
 T ∂V (V – b)T (V – b)2
∂M = ∂N , dP is an exact differential.
Since  ∂T  ∂V
 V  T
∂M =  ∂  ∂P   = ∂2P
 ∂T  ∂T ∂V 
 V  TV
∂T ∂V

∂N =  ∂ ∂P  = ∂2P


∂V  
 T ∂V ∂TVT ∂V ∂T
∂2P ∂2P
i.e., =
∂T ∂V ∂V ∂T
CYCLIC RULE
Let z be a function of x, y such that z = f1 (x, y)


∂z ∂z
dz =   dx +   dy ... (1.17)
∂xy ∂yx
Since z is a function of x, y; x is also a function of z, y such that x = f2 (z, y)


∂x ∂x
dx =   dz +   dy ... (1.18)
∂zy ∂yz
Substituting for dx from Equation (1.18) into Equation (1.17) yields
∂z ∂x ∂x  ∂z
dz =     dz +   dy +   dy
∂xy ∂zy ∂yz  ∂yx

∂z ∂x dz + ∂z ∂x + ∂z  dy


dz =   ∂z ∂x ∂y ∂y 
∂xy  y  y z x

∂z ∂x = 1
But ∂x ∂z
 y  y
With this, the above equation becomes

∂z ∂x ∂z 


dz = dz +     +    dy
∂xy ∂yz ∂yx
∂z ∂x ∂z 
∴ 0 =     +    dy
∂xy ∂yz ∂yx
Chemical Engineering Thermodynamics - I 1.35 Introduction and Basic Concepts

dy ≠ 0 as it represents arbitrary change in y, so that


∂z ∂x + ∂z = 0
∂x ∂y  
 y  z ∂yx
∂z ∂x = – ∂z
∂x ∂y ∂y
 y  z  x
∂z ∂x
∂x ∂y
 y  z
∴ = –1
∂z
∂y
 x
∂z = 1
But ∂y
 x ∂y
∂z
 x
∂z ∂x ∂y = – 1
∴ ∂x ∂y ∂z ... (1.19)
 y  z  x
Equation (1.19) is called the cyclic rule or cyclic relation. It is applicable only in case of
state functions.
∂z ∂x = 1
∂x ∂z
 y  y
∂z = 1
⇒ ∂x ... (1.20)
 y ∂x
∂z
 y
Equation (1.20) is known as the reciprocal relation.
(5) Verify the reciprocal relation for 1 mole of an ideal gas :
For 1 mole of an ideal gas,
PV = RT ⇒ zx = Ry
RT
∴ P = V
Differentiating P partially with respect to V yields
 ∂P  = –RT
∂V
 T V2
RT
We have, V = P
Differentiating V partially with respect to P (i.e., differentiating V with respect to P, keeping
T constant) gives
∂V = –RT
 ∂P 
 T P2
Chemical Engineering Thermodynamics - I 1.36 Introduction and Basic Concepts

 ∂P  ∂V = –RT × – RT = (RT)2 = (PV)2 = 1


∴ ∂V  ∂P 
 T  T V2 P2 (PV)2 (PV)2

 ∂P  ∂V = 1 ⇒  ∂P  = 1
∂V  ∂P  ∂V
 T  T  T ∂V
 ∂P 
 T
∂z ∂x ∂z = 1
i.e.,     = 1 ⇒ ∂x
∂xy ∂zy  y ∂x
∂z
 y
 ∂P  = –RT and ∂V = – RT
∂V  ∂P 
 T V2  T P2

 ∂P  ∂V = – RT × –RT = (RT)2 = (RT)2 = 1


∂V  ∂P 
 T  T V2 P2 (PV)2 (RT)2

(6) Verify the cyclic rule for 1 mole of an ideal gas.


For 1 mole of an ideal gas, we have
PV = RT
RT
∴ P = V
Differentiating P w.r.t. T keeping V constant, we get
∂P = R
∂T
 V V
RT
We have, V = P
Differentiating V w.r.t. P keeping T constant gives
∂V = –RT
 ∂P 
 T P2
PV
We have, T = R
Differentiating T w.r.t. V keeping P constant yields
 ∂T  = P
∂V
 P R
∂P  ∂T  ∂V = R × P × –RT = –RT = – RT = – 1
∴ ∂T ∂V  ∂P 
 V  P  T V R P2 PV RT
∂P  ∂T  ∂V = –1 which is the cyclic rule/cyclic relation
∴ ∂T ∂V  ∂P 
 V  P  T
RT
PV = RT ⇒ P = V
Chemical Engineering Thermodynamics - I 1.37 Introduction and Basic Concepts

 ∂P  = – RT
∴ ∂V ... Differentiation of P w.r.t. V keeping T constant
 T V2
RT
V = P

∂V = R
∴  ∂T  ... Differentiation of V w.r.t. T keeping P constant
 P P
PV
T = R

∂T = V
∴ ∂P ... Differentiation of T w.r.t. P keeping V constant
 V R

 ∂P  ∂V ∂T = –RT × R × V = – RT = – RT = –1


∴ ∂V  ∂T  ∂P
 T  P  V V2 P R PV RT

 ∂P  ∂V ∂T = –1 ... which is the cyclic rule


∴ ∂V  ∂T  ∂P
 T  P  V

(7) The van der Waals equation of state P + V2 (V – b) = RT is pressure explicit
a
  
[i.e., volume and temperature appear as independent variables] and not volume explicit.
Let us find   for a van der Waals gas using the cyclic rule.
∂V
∂TP
∂V ∂T  ∂P  = – 1
 ∂T  ∂P ∂V
 P  V  T
∂V = –1
 ∂T 
 P ∂T  ∂P 
∂P ∂V
 V  T
∂T = 1
But ∂P … reciprocal relation
 V ∂P
∂T
 V
∂P
– 
∂V = ∂TV
∴  ∂T 
 P  ∂P 
∂V
 T
The van der Waals equation of state is
P + a  (V – b) = RT
 V 2
RT a
Rearranging, P = V – b – V2 ... (1.21)
Chemical Engineering Thermodynamics - I 1.38 Introduction and Basic Concepts

Differentiating Equation (1.21) with respect to T keeping V constant yields


∂P = R – 0 = R
∂T
 V V–b V–b

Differentiating Equation (1.21) with respect to V keeping T constant yields


 ∂P  = –RT + 2a
∂V
 T (V – b)2 V3
∂P R –R
–  – (V – b)
∂V = ∂TV (V – b)
∴  ∂T  = 2a =
 P  ∂P  –RT 2a RT
∂V +
(V – b) V3
2 V3 – (V – b)2
 T
Let f be a function of x, y and z and there exists a relation between x, y and z so that f is a
function of any two of x, y and z and any one of x, y and z is regarded as a function of f and any
one of x, y and z. The chain rule of partial differentiation is
∂x = ∂x ∂y
∂z    
 f ∂yf ∂zf
∂x ∂y ∂z
or ∂y ∂z ∂x = 1 ... (1.22)
 f  f  f
The chain rule of partial differentiation is true for any number of changes or transformation
as long as the same variable (say f) is held constant.
The chain rule of partial differentiation, the cyclic relation, the reciprocal relations are
extensively used to derive the thermodynamic property relations [chapter-5]. The
relationship between thermodynamic properties can also be derived using the method of
Jacobians. Jacobians are determinants, with partial derivatives as their elements.
If x and y are differentiable functions of independent variables u and v
x = f2 (u, v) and y = f2 (u, v)
[x‚ y]
then the Jacobian of x, y with respect to u, v, denoted by J [x, y / u, v] or simply [u‚ v] for
convenience, is defined as

∂x ∂y
∂u ∂u
[x‚ y]  v  v
[u‚ v] = ∂x ∂y
∂v ∂v
 u  u

∂x ∂y ∂x ∂y


=     –    ... (1.23)
∂uv ∂vu ∂vu ∂uv
Chemical Engineering Thermodynamics - I 1.39 Introduction and Basic Concepts

PROPERTIES OF JACOBIANS
• [x, x] = 0 ... (1.24)
• [x, y] = – [y, x] ... (1.25)
[x‚ z] ∂x
=   ... (1.26)
• [y‚ z] ∂yz
Let us verify this property.
[y‚ z] ∂y
=  
[x‚ z] ∂xz

∂y ∂z ∂y 0


∂x ∂x ∂x
[y‚ z]  z  z  z
[x‚ z] = =
∂y ∂z ∂y 1
∂z ∂z ∂z
 x  x  x

∂y ∂y
=   ×1–  ×0
∂xz ∂zx
[y‚ z] ∂y
∴ =  
[x‚ z] ∂xz
[x‚ y] [x‚ y] / [s‚ t]
• [u‚ v] = [u‚ v] / [s‚ t] ... (1.27)

[x‚ y] [u‚ v] [x‚ y]


• [u‚ v] · [m‚ n] = [m‚ n] ... (1.28)

• The exact differential equation,


∂z ∂z
dz =   dx +   dy = M dx + N dy
∂xy ∂yx
in terms of Jacobians can be written as
[z‚ y] [z‚ x]
dz = [x‚ y] dx + [y‚ x] dy ... (1.29)
This can be written as
[z‚ y] [z‚ x]
dz = [x‚ y] dx – [x‚ y] dy
[x, y] dz = [z, y] dx – [z, x] dy
[x, y] dz + [y, z] dx + [z, x] dy = 0 ... (1.30)
This equation is similar to a cyclic relation among the variables x, y and z.
• The exact differential, dz = M dx + N dy in terms of Jacobians can be written as
[z, u] = M [x, u] + N [y, u] ... (1.31)
where u is any dummy variable.
Chemical Engineering Thermodynamics - I 1.40 Introduction and Basic Concepts

(7) Let us verify this property.


dz = M dx + N dy
Dividing this equation by v keeping u constant gives
∂z = M ∂x + N ∂y
∂v ∂v ∂v
 u  u  u
Eliminating the partial derivatives in favour of the Jacobians gives
∂z = [z‚ u]
∂v
 u [v‚ u]

∂x = [x‚ u]
∂v
 u [v‚ u]

∂y [y‚ u]
and ∂v = [v‚ u]
 u
[z‚ u] [x‚ u] [y‚ u]
∴ [v‚ u] = M [v‚ u] + N [v‚ u]

Multiplying the above equation by [v, u], we get


[z, u] = M [x, u] + N [y, u] ... thus verified Equation (1.31)
The differential forms of the fundamental relations can be written in terms of Jacobians
as
dU = T dS – P dV ⇒ [U, x] = T [S, x] – P [V, x] ... (1.32)
dH = T dS + V dP ⇒ [H, x] = T [S, x] + V [P, x] ... (1.33)
dG = – S dT + V dP ⇒ [G, x] = – S [T, x] + V [P, x] ... (1.34)
dA = – S dT – P dV ⇒ [A, x] = – S [T, x] – P [V, x] … (1.35)
where x is any dummy variable.

SOLVED EXAMPLES
Example 1.1 :
A man weighs 600 N on the earth’s surface where the gravitational acceleration is 9.81 m/s2.
Calculate the weight of the man on the moon where the gravitational acceleration is 1.67 m/s2.
Solution :
The gravitational force acting on a body of mass m kg is given by
F = mg ... Force when a mass (m) acts under gravitational acceleration
The force of gravity on a body is usually called the weight of the body and hence the unit of
weight is newton.
Chemical Engineering Thermodynamics - I 1.41 Introduction and Basic Concepts

The force of gravity on the man on the earth’s surface is given by


F = mg , F = 600 N, g = 9.81 m/s2
600 = m (9.81)
m = 61.16 kg ... mass of the man
The mass of the man remains constant whether it is on the earth’s surface or on the moon,
but the weight or the force of gravity on the man depends upon the location, as the acceleration
due to gravity is different at different locations.
The force of gravity acting on the man on the moon’s surface is given by
F on the moon = mg where g = 1.67 m/s2
= 61.16 × 1.67
= 102.14 N
The weight of the man on the moon = 102.14 N ... Ans.
As m is constant at any location, we can write
F = F
gon the earth gon the moon

F, force on the moon = g


F

 on the earth × g on the moon
600
= 9.81 × 1.67 = 102.14 N ... Ans.

Example 1.2 :
If a man circling the earth in a spaceship weighs 300 N, where the local gravitational
acceleration is 3.35 m/s2, what would be the mass of the man and his weight on the earth, where
the gravitational acceleration is 9.81 m/s2 ?
Solution :
The gravitational force acting on a body of mass m kg is given by
F = mg
(This force of gravity on a body is usually called the weight of the body.)
Case I : F = 300 N, g = 3.35 m/s2
∴ 300 = m × 3.35
∴ m = 89.55 kg
The mass of the man is 89.55 kg.
The mass of the man remains constant, i.e., it remains the same irrespective of the location.
The mass of a body does not change with location.
∴ The mass of the man on the earth = 89.55 kg ... Ans.
Chemical Engineering Thermodynamics - I 1.42 Introduction and Basic Concepts

The weight of a man on the earth is the force acting on the man on the earth and is given by
F on the earth = mg
where m = 89.55 kg and g = 9.81 m/s2
∴ F on the earth = 89.55 × 9.81
= 878.4855 ≈ 878.5 N
The weight of the man on the earth = 878.5 N ... Ans.
Example 1.3 :
A mercury manometer is used to measure pressure inside a vessel. One end of the
manometer is open to the atmosphere and the manometer reads 400 mm (manometer reading).
The atmospheric pressure is 1.01325 bar. Find the absolute pressure prevailing in the vessel.
Data : ρ of mercury = 13.56 kg/m3 and g = 9.81 m/s2
Solution :
Refer to Fig. E 1.3. From the figure, we can write
P1 = P2 + hρg
where h is the difference in the height of the mercury columns (manometer reading)
h = 400 mm = 0.40 m
P2 = Atmospheric pressure = 1.01325 bar = 1.01325 × 105 N/m2
P1 = Pressure prevailing in the vessel in N/m2 = ?
ρ = Density of manometric fluid = 13.56 kg/m3
g = Acceleration due to gravity = 9.81 m/s2
[Note that 1 N = 1 (kg·m)/s2]

To vessel Open to atmosphere


(P1)

Fig. E 1.3
∴ P1 = 1.01325 × 105 + 0.40 × 13.56 × 103 × 9.81
= 154534.44 N/m2
= 1.545 × 105 N/m2
= 1.545 × 105 Pa = 1.545 × 102 kPa
The absolute pressure prevailing in the vessel is 1.545 × 105 N/m2 = 1.545 × 102 kPa ... Ans.
Chemical Engineering Thermodynamics - I 1.43 Introduction and Basic Concepts

Example 1.4 :
The potential energy of a body of mass 20 kg is 3.5 kJ. Find the height of the body from the
ground. If the same body is moving with a velocity of 50 m/s, find its kinetic energy.
Solution :
The potential energy of a body is given by
P.E. = mgz
where m = 20 kg, P.E. = 3.5 kJ = 3500 J, g = 9.81 m/s2
z = Height / elevation of the body from the ground
∴ 3500 = 20 × 9.81 × z
∴ z = 17.84 m
Height of the body from the ground = 17.84 m ... Ans.
The kinetic energy of a body is given by
1
K.E. = 2 mu2

where m = 20 kg and u = 50 m/s


1
∴ K.E. = 2 × 20 × (50)2 = 25000 (kg·m2)/s2

kg.m
= 25000 s2 · m = 25000 N·m
= 25000 J = 25 kJ
Kinetic energy of the body = 25 kJ ... Ans.
Example 1.5 :
A car having a mass of 1200 kg is running at a speed of 60 km/h. Calculate the kinetic
energy of the car in kJ. Also calculate the work to be done on the car to bring it to a stop.
Solution :
m = 1200 kg, u = 60 km/h = 60000 m/h = 16.667 m/s
1
K.E. of the car = 2 mu2

1
= 2 × (1200) × (16.667)2

= 166673 kg (m2/s2) ≡ N·m ≡ J


= 166673 J
= 1.66673 × 105 J ≈ 1.67 × 105 J
= 1.67 × 102 kJ ... Ans.
Chemical Engineering Thermodynamics - I 1.44 Introduction and Basic Concepts

Work to be done on the car to bring it to a stop :


u1 = 16.667 m/s, u2 = 0 m/s ... car at rest
1 2 1 2
Work to be done = 2 mu1 – 2 mu2
1 1
= 2 × (1200) × (16.67)2 – 2 × (1200) × (0)2
= 1.66733 × 105 J
≈ 1.67 × 102 kJ ... Ans.
Example 1.6 :
A man weighing 700 N takes 2.5 min to climb up a staircase. Find the power developed in
the man, if the staircase is made of 20 stairs, each of 0.18 m height.
Solution :
Total height to climb up the staircase = 20 × 0.18 = 3.6 m
Displacement by the man = 3.6 m
Work done in climbing = Force × Displacement
 up the staircase 
= 700 × 3.6
= 2520 N·m = 2520 J
Work done
Power developed = Time , Time = 2.5 min
2520
=
2.50 × 60
= 16.8 J/s = 16.8 W ... Ans.
Example 1.7 :
Steam supplied to an engine liberates 5000 J of heat. The efficiency of the engine is 40%.
Find the height to which a body of mass 10 kg can be lifted using the work output from the
engine.
Solution :
Heat liberated from the steam and received by the engine = 5000 J
Efficiency of the engine = 40% or 0.40
Work output (W)
Thermal efficiency of the engine = Heat received by the engine (Q)
∴ Work output from the engine = 0.40 × (5000) = 2000 J
This work will be utilised to lift a body of mass 10 kg.
Work output = mgz (potential energy)
∴ 2000 = 10 × 9.81 × z
2000
∴ z = = 20.387 ≈ 20.39 m
10 × 9.81
Height to which the body can be lifted = 20.39 m ... Ans.
Chemical Engineering Thermodynamics - I 1.45 Introduction and Basic Concepts

Example 1.8 :
Nitrogen gas is confined in a cylinder and its pressure is maintained by a weight placed on
the piston. The mass of the piston and the weight together is 100 kg. Assuming frictionless piston,
find :
(i) the force exerted by the atmosphere, the piston and the weight on the gas,
(ii) the pressure of the gas,
(iii) the work done by the gas (in kJ) if the gas is allowed to expand pushing up the piston
and the weight by 500 mm.
Data : Piston diameter = 200 mm, g = 9.81 m/s2, Atmospheric pressure = 1.01325 bar
Solution :
Piston diameter = 200 mm = 0.20 m
π π
Area of the piston = 4 (D2) = 4 (0.20)2 = 0.0314 m2

Mass of the piston + Weight on it = 100 kg


Force exerted by the piston + Weight = F1 = mg = 100 × 9.81 = 981 N
Atmospheric pressure = 1.01325 bar = 1.01325 × 105 N/m2
Force exerted by the atmosphere = Pressure × Area of the piston
F2 = 1.01325 × 105 × 0.0314
= 3181 N
Total force acting on the gas = F1 + F2
= 981 + 3181
= 4162 N = 4.16 kN ... Ans. (i)
Force
Pressure of the gas = Area

4162
= 0.0314

= 132547.8 N/m2 ≈ 1.325 × 105 N/m2


= 1.325 bar ... Ans. (ii)
Displacement of the piston due to expansion = 500 mm = 0.50 m
Work done by the gas = Force × Displacement of the piston
= 4162 × 0.50
= 2081 N·m = 2081 J ≈ 2.081 kJ ... Ans. (iii)
Chemical Engineering Thermodynamics - I 1.46 Introduction and Basic Concepts

Example 1.9 :
A balloon that was originally empty is being filled with hydrogen from a cylinder. The
atmospheric pressure is 1.01325 bar. Estimate the work done by the balloon-cylinder system
when the balloon attains a spherical shape 6 m in diameter.
Solution :
Balloon diameter = 6 m
Here work is done by the balloon-cylinder system against the atmospheric pressure.
Atmospheric pressure = 1.01325 bar = 1.01325 × 105 N/m2
Initial volume of the balloon = 0 m3 ... empty volume
4 π
Final volume of the balloon = 3 πR3 = 6 D3, where D = 6 m

π
= 6 × (6)3

= 113.097 m3
 Work done by the  = P ∆V
balloon-cylinder system
= 1.01325 × 105 × 113.097
= 11459553 N·m
= 1.1459553 × 107 N·m or J
= 1.1459553 × 104 kJ
≈ 1.146 × 104 kJ ... Ans.
Example 1.10 :
A piston encloses a gas within a cylinder. The
piston is restrained by a linear spring as shown in
Fig. E 1.10 (a). The initial volume and pressure are
0.001 m3 and 1.5 bar (150 kPa) respectively. At the
initial position/condition, the spring touches the
piston, but exerts no force. The gas is heated until the
volume becomes three times the original volume and Gas
the pressure rises to 10 bar (1000 kPa).
Fig. E 1.10 (a)
(i) Draw the P-V diagram for the process.
(ii) Calculate the work done by the gas.
(iii) Calculate the work done against the piston and against the spring.
The force relation for a linear spring is :
Fspring = ks d
where ks is the spring constant and d is the distance displaced from the position of zero force.
Chemical Engineering Thermodynamics - I 1.47 Introduction and Basic Concepts

Solution :
The pressure exerted on the gas by the piston spring assembly is the pressure exerted by the
gas on the piston-spring assembly (as the process is in quasi-equilibrium state).
Gas pressure = Pressure exerted by the gas = P
P = Ppiston + Pspring
Fspring
P = Ppiston + A , where A is the area of piston

ks d
= Ppiston + A … (1)

where d is the distance displaced from the position of zero force, i.e., from the volume
V1 = 0.001 m3
∴ d = 0 at V1 and d = d at V ... new volume
V – V1 = A (d – 0)
V – V1
∴ d = A ... distance displaced in terms of the volume

Initially the spring exerts no force on the gas.


ks (V – V1)
That is, F = ks d = A = 0, when V = V1

Substituting for ks d in Equation (1) gives


ks
P = Ppiston + A2 (V – V1)

1000 kPa

Work done against


P 2 spring (triangular area)

150 kPa
Work done against
piston (rectangular area)
1
0,0 0.001 0.002 0.003
V
Fig. E 1.10 (b) : P-V diagram for the process
Chemical Engineering Thermodynamics - I 1.48 Introduction and Basic Concepts

ks
The above equation is a straight line on a P-V diagram with a slope equal to A2 .

ks
P = Ppiston + A2 (V – V1)

At P = 150 kPa, V = V1 = 0.001 m3


ks
∴ 150 = Ppiston + A2 (0)

∴ Ppiston = 150 kPa


At P = 1000 kPa, V = 3V1 = 3 × 0.001 = 0.003 m3 … mentioned in the statement
ks
∴ 1000 = 150 + A2 (0.003 – 0.001)

ks
∴ 3
A2 = 425000 kPa/m = 425000 kN/m
5

Therefore, the pressure-volume relationship is


P = 150 + 425000 (V – 0.001) ... kPa
where V is in m3.
2

Total work done by the gas = ⌠


⌡ P dV
1
2

W = ⌠
⌡ [150 + 425000 (V – 0.001)] dV
1
0.003

= ⌠
⌡ [150 + 425000 V – 425] dV
0.001
0.003
= 425000 2 – 275 V
V 2

 0.001
425000
= 2 [(0.003)2 – (0.001)2] – 275 (0.003 – 0.001)
= 1.15 kN·m = 1.15 kJ ... Ans. (i)
The work done by the gas to raise the piston without the spring is given by
2

Wp = ⌠
⌡ Ppiston dV
1
2 0.003
0.003
= ⌠
⌡ 150 dV = 150 ⌠
⌡ dV = 150 [V]0.001
1 0.001

= 150 (0.003 – 0.001) = 0.3 kN·m = 0.30 kJ ... Ans. (ii)


Chemical Engineering Thermodynamics - I 1.49 Introduction and Basic Concepts

The work done by the gas against the spring is given by


2

Ws = ⌠
⌡ Pspring dV
1
2
ks
= ⌠
⌡ A2 (V – V1) dV
1
2
ks
= A2 ⌠
⌡ (V – 0.001) dV
1
0.003
= 425000  2 – 0.001 V
V2
 0.001
1 
= 425000 2 [(0.003)2 – (0.001)2] – 0.001 (0.003 – 0.001)
 
= 0.85 kN·m = 0.85 kJ ... Ans. (iii)
Total work done = Work done to raise piston + Work done against spring
 by the gas 
= 0.3 kJ + 0.85 kJ = 1.15 kJ
Example 1.11
A constant volume helium gas thermometer registers a pressure of 1000 mm Hg and
1366 mm Hg at ice point and steam point, respectively. The thermometer follows the relation :
T = A + BP
where T is in °C and P is in mm Hg.
Find the atmospheric temperature, if the thermometer registers the reading of 1100 mm Hg
when left standing in atmosphere.
Solution :
T = A + BP
Let us evaluate A and B.
At ice point, T = 0°C, P = 1000 mm Hg
At steam point, T = 100°C, P = 1366 mm Hg
∴ 0 = A + 1000 B … (1)
and 100 = A + 1366 B … (2)
366 B = 100 ∴ B = 0.273
100 = A + 1366 (0.273)
∴ A = – 272.918 = – 273
Chemical Engineering Thermodynamics - I 1.50 Introduction and Basic Concepts

The relationship between T and P is


T = – 273 + 0.273 P = 0.273 P – 273

T = 273 1000 – 1
P
i.e.,
 
To find T (atmospheric temperature) when P = 1100 mm Hg

T = 273 1000 – 1
1100
 
= 27.3°C … Ans.
Example 1.12 :
The relation T = A ln X + B is used to establish the temperature scale of a certain
thermometer, where A and B are constants and X is the thermometric property of the fluid in the
thermometer. The thermometric property is found to be 1.5 at the ice point and 7.5 at the steam
point. Determine the temperature corresponding to the thermometric property of 3.5 on the
Celsius scale.
Solution :
On the Celsius scale :
Ice point = 0°C and steam point = 100°C
We have, T = A ln X + B
At T = 0°C, X = 1.5
∴ 0 = A ln 1.5 + B
0.4054 A + B = 0 … (1)
At T = 100°C, X = 7.5
∴ 100 = A ln (7.5) + B
2.015 A + B = 100 … (2)
Subtracting (1) from (2) gives
A = 62.127
and B = – 25.185
To determine T at X = 3.5
T = A ln X + B
T = 62.127 ln X – 25.185
T = 62.127 ln (3.5) – 25.185
= 52.64°C … Ans.
Example 1.13 :
The following relation is used to establish the temperature scale of a thermometer.
T = A ln X + B, where A and B are constants.
The thermometer is calibrated using the ice and steam points as fixed reference points.
Chemical Engineering Thermodynamics - I 1.51 Introduction and Basic Concepts

Show that the unknown temperature T (in °C) is given by


 ln (X/Xi) 
T = 100 ln (X /X )
 s i 
where Xi and Xs are the values of the thermometric property X of the thermometer at the ice
point and steam point, respectively.
Solution :
The relationship between T and X is
T = A ln X + B … (1)
At ice point : T = 0°C and X = Xi
∴ 0 = A ln Xi + B … (2)
At steam point :
T = 100°C and X = Xs
100 = A ln Xs + B … (3)
Subtracting (2) from (3) gives
A ln Xs – A ln Xi = 100
A [ln Xs – ln Xi] = A ln (Xs/Xi) = 100
100
∴ A = ln (X /X )
s i
Substituting the value of A in (2) gives
100
0 = ln (X /X ) (ln Xi) + B
s i
–100 ln Xi
∴ B = ln (X /X )
s i
Substituting the values of A and B in (1), we get
100 ln X 100 ln Xi
T = ln (X /X ) – ln (X /X )
s i s i
100
T = ln (X /X ) [ln X – ln Xi]
s i
100 ln (X/Xi)
T = ln (Xs/Xi) … Ans.

Example 1.14 :
1 kg of a certain gas is compressed reversibly according to a law PV = 0.25 where P is bar
and V in m3/kg. The final value is 1/4th of the initial volume. Find the work done on the gas.
Solution :
The work done is given by
2
W = ⌠
⌡ P dV
1
Chemical Engineering Thermodynamics - I 1.52 Introduction and Basic Concepts

We have : PV = 0.25, P is in bar and V is in m3/kg.


0.25
∴ P = V , bar
0.25 × 105
= V , N/m2 or Pa [since 1 bar = 105 Pa]
Substituting for P, we get
2
0.25 × 105
W = ⌠
⌡ V dV = 0.25 × 105 ln (V2/V1)
1

= 0.25 × 105 ln  V  = 0.25 × 105 ln (0.25)


1/4 · V1
 1 
since V2 = 1/4 · V1 … given
N m3
= – 34657.35 m2 × kg … units of P × units of dV
= – 34657.35 N.m/kg
W in N.m = m in kg × W in N.m/kg
= 1 × (– 34657.35), [since m of the gas is 1 kg]
W = – 34657.35 N.m
The minus sign indicates that the work is done on the gas (the system). Therefore, by
convention.
Work done on the gas = – W = 34657.35 N.m = 34657.35 J … Ans.
Example 1.15 :
A fluid at a pressure of 300 kPa and with a specific volume of 0.18 m3/kg contained in a
cylinder behind a piston expands reversibly to a pressure of 60 kPa, according to the following
law :
c
P = V2 where c is a constant and V is volume in m3/kg.

Calculate the work done by the fluid.


Solution :

1
P1
P1 = 300 kPa
P2 = 60 kPa
P
P2 2

V
Fig. E 1.15
Chemical Engineering Thermodynamics - I 1.53 Introduction and Basic Concepts

We have : P1 = 300 kPa = 3 × 105 Pa


P2 = 60 kPa = 0.60 × 105 Pa
V1 = 0.18 m3/kg
2
Work done by the fluid = shaded area = ⌠
⌡ P dV
1
2
W = ⌠
⌡ P dV
1

c
Given : P = V2
2 2 V
V 
c –2 + 1
∴ W = ⌠
⌡ V2 dV = c
–2 + 1V1
1

W = c V – V 
1 1
 1 2

c
We have : P = V2
2
∴ c = PV2 = P1V1 = 3 × 105 × (0.18)2
= 9720 Pa . (m3/kg)2
c
We have : P = V2

V = P
c 1/2
 

V2 = P  = 
c 1/2 9720 1/2
∴ = 0.402 m3/kg
 2 0.60 × 105

Work done by the fluid, W = c V – V  = 9720 0.18 – 0.402


1 1 1 1
 1 2  
m6 kg
= 29821 Pa · kg2 × m3

N m3
= 29821 Pa · m3/kg = 29821 m2 × kg

= 29821 (N.m)/kg
= 29821 J/kg (29.821 kJ/kg) … Ans.
Example 1.16 :
The work supplied to a closed system having an initial volume of 0.80 m3 is 160 kJ. The
pressure of the system changes according to P = 7 – 3V, where P is in bar and V is in m3. Find
the final volume and pressure of the system.
Chemical Engineering Thermodynamics - I 1.54 Introduction and Basic Concepts

Solution :
Work supplied to the closed system = Work done on the closed system
Work done on the system, W = – 160 kJ … by convention
Initial volume = V1 = 0.80 m3
Pressure-volume relationship : P = 7 – 3 V, P in bar and V in m3
The work done on the system is given by
2
W = ⌠
⌡ P dV
1
2
– 160 = ⌠
⌡ P dV … in kJ
1
2
– 160 × 103 = ⌠
⌡ P dV … in J … (1)
1

The P-V relationship is


P = 7 – 3 V, where P is in bar and V in m3
But P in Equation (1) is in Pa, i.e., N/m2.

P = (7 – 3 V) × 105 N/m2, P = (7 – 3 V) bar × 1 bar 


105 N/m2

 
Substituting this P in Equation (1), it becomes
2 2
– 160 × 103 = ⌠
⌡ (7 – 3 V) × 10 dV = 10 ⌠
5 5
⌡ (7 – 3V) dV
1 1

V2 V2

– 160 × 103 = 105 ⌠ (7 – 3 V) dV =


⌡ 105 ⌠ (7 – 3V) dV

V1 0.80

V2
– 1.60 = 7V – 2 V2
3
 0.80
2
– 1.60 = 7 (V2 – 0.80) – 1.5 [V2 – (0.80)2]
2
= 7V2 – 1.5V2 – 4.64 … in J
2
1.5 V2 – 7 V2 + 3.04 = 0
– (–7) ± [(–7)2 – 4 (1.5) (3.04)]1/2 –b ± (b2 – 4ac)1/2
V2 = , 2a
2 × (1.5)
7 ± 5.55
= 3 = 4.183 or 0.483 m3
Chemical Engineering Thermodynamics - I 1.55 Introduction and Basic Concepts

The root (4.183 m3) is incompatible with the given problem (since the work is done on the
system so that its final volume must be less than the initial volume (0.80 m3). Therefore, the
correct root is 0.483 m3.
Final volume of the system = V2 = 0.483 m3 … Ans.
Final pressure of the system = P2 = 7 – 3 V2
= 7 – 3 × 0.483 = 5.551 bar = 555.1 kPa … Ans.
Example 1.17 :
1 kg of a fluid at a pressure of 20 bar containing in a cylinder is allowed to expand behind a
piston according to a law PV2 = constant until the volume is doubled. The fluid is then cooled
reversibly at constant pressure until the piston comes to its original position. Afterwards, heat is
supplied reversibly with the piston firmly locked in position till the pressure rises to the original
pressure (20 bar). For an initial volume of 0.05 m3, calculate the net work done by the fluid.
Solution :

Fig. E 1.17
We have : Mass of fluid = 1 kg

Initial volume = V1 = 0.05 m3

Final volume = V2 = 0.10 m3 [since V2 = 2V1 … given]

Initial pressure = P1 = 20 bar = 20 × 105 Pa = 20 × 105 N/m2

For the process 1-2 :


2 2
PV2 = c ∴ P1V1 = P2V2
2 2
∴ P2 = P1 V1 / V2 = 20 × 105 × (0.05)2/(0.1)2

= 5 × 105 Pa = 5 × 105 N/m2 = 5 bar


Chemical Engineering Thermodynamics - I 1.56 Introduction and Basic Concepts

2
⌠ P dV
Work done by the fluid from 1 to 2 = Area 1 – 2 – 5 – 4 – 1 = ⌡
1
2 2
W1–2 = ⌠
c 1 1 
⌡ P dV = ⌠
⌡ V2 dV = c V1 – V2 
1 1
2
We have : PV2 = c ∴ c = P1V1

W1–2 = P1V1 V1 – V  = 20 × 105 × (0.05)2 0.05 – 0.10


2 1 1 1 1

 2  
= 50000 N.m
or W1–2 = 50000 N.m by 1 kg
For the process 2 to 3 :
While cooling the fluid at constant pressure [P = P2 = 5 × 105 Pa] to its original volume, the
work is done on the fluid.
The work done on the fluid from 2 to 3 is
–W2–3 = Area 2 – 3 – 4 – 5 – 2 = P2 (V2 – V1)
= 5 × 105 [0.10 – 0.05], (N/m2) × m3
= 25000 N.m
–W2–3 = 25000 N.m by 1 kg
For the process 3-1 :
The work done during the process 3 -1 is
W3–1 = 0
[Since the piston is firmly locked in position, the volume remains constant and addition of
heat increases pressure. It is a constant volume process and for the constant volume process,
we know that W = 0, i.e., there is no displacement work since dV = 0.]
Net work done by the fluid,
W = Wnet = enclosed area 1 – 2 – 3 – 1
= W1–2 + W2–3 + W3–1
= 50000 – 25000 + 0
= 25000 N.m = 25000 kJ = 25 kJ … Ans.
[Since –W2–3 = 25000 N.m, W2–3 = – 25000 N.m]
Example 1.18 :
A system delivers 50 kJ of work during reversible non-flow process. The pressure of the
system varies as
12.8
P = V – V2, where P is in bar and V is in m3
Determine P2 and V2, if P1= 3.5 bar.
Chemical Engineering Thermodynamics - I 1.57 Introduction and Basic Concepts

Solution :
Work done by the system = Work delivered by the system
= 50 kJ = 50 × 103 J = 5 × 104 J = 5 × 104 N.m
The work done is given by
2
W = ⌠
⌡ P dV … in J or N.m
1
2
5 × 104 = ⌠
⌡ P dV … in J or N.m
1

Given : P1 = 3.5 bar = 3.5 × 105 Pa= 3.5 × 105 N/m2


12.8
P = V – V2, in bar where V is in m3

P =  V – V2 × 105, in N/m2 or Pa


12.8
 
2
5 × 104 = ⌠ 12.8 2
⌡  V – V  × 10 dV
5

1
1 3 3
0.50 = 12.8 ln (V2/V1) – 3 [V2 – V1 ] … (1)
Let us find the initial volume from the initial pressure.
12.8
We have : P = V – V2 , where P is in bar and V in m3
12.8 2
P1 = V – V1
1

Given : P1 = initial pressure = 3.5 bar


12.8 2
∴ 3.5 = V – V1
1

Find V1 by a trial and error procedure. Assume a value of V1, evaluate R.H.S. and check
RHS ≈ LHS or not. If not, repeat the procedure.
Trial - 1 : V1 = 2 m3 : LHS = 3.5 and RHS = 2.4
Trial - 2 : V1 = 1.75 m3 : LHS = 3.5 and RHS = 4.2518
Trial - 3 : V1 = 1.85 m3 : LHS = 3.5 and RHS = 3.4964
∴ V1 = 1.85 m3 since for this value LHS (3.5) ≈ RHS (3.4964)
Substituting V1 = 1.85 m3, Equation (1) becomes
3
0.50 = 12.8 ln V2 – 12.8 ln (1.85) – 0.333 V2 + 0.333 (1.85)3
3
0.50 = 12.8 ln V2 – 7.874 – 0.333 V2 + 2.108
3
6.266 = 12.8 ln V2 – 0.333 V2
Adopt a trial and error procedure to find V2.
Since the system delivers/performs work, V2 has to be greater than V1 (1.85 m2).
Chemical Engineering Thermodynamics - I 1.58 Introduction and Basic Concepts

Trial - 1 : V2 = 2 m3
RHS = 12.8 ln (2) – 0.333 (2)3
= 6.2083 ≠ LHS (6.266)
Trial - 2 : V2 = 2.1 m3, RHS = 6.4128
Trial - 3 : V2 = 2.05 m3, RHS = 6.3195
Trial - 4 : V2 = 2.03 m3, RHS = 6.2771
Trial - 5 : V2 = 2.027 m3, RHS = 6.2706
Trial - 6 : V2 = 2.025 m3, RHS = 6.2662
∴ RHS (6.2662) ≈ LHS (6.266) for V2 = 2.025 m3
∴ The final volume (V2) is 2.025 m3 … Ans.
12.8 2
We have : P = V – V , bar
12.8 2
P2 = V – V2
2

12.8
= 2.025 – (2.025)2

= 2.22 bar
The final pressure (P2) is 2.22 bar. … Ans.

❐❐❐
Chapter ... 2
FIRST LAW OF THERMODYNAMICS

The two modes of energy transfer between a system and its surroundings are heat and work.
The first law of thermodynamics is essentially the law of conservation of energy with
particular reference to heat and work.
The first law of thermodynamics states that energy can neither be created nor destroyed
during a process, although it may be converted from one form to another.
The first law can also be stated in the following alternative forms / ways :
• Whenever a certain quantity of one kind energy disappears, an exactly equivalent
amount of some other kind (kinds) must appear.
The total energy of an isolated system must remain constant.
Since the combination of a system and its surroundings may be referred to as an isolated
system, the above statement signifies that any loss or gain of energy by a system must
be exactly equivalent to the gain or loss of energy by the surroundings.
• Since the universe can be treated as an isolated system, the energy of the universe is
conserved.
• It is impossible to construct a perpetual motion machine of the first kind. It is
impossible to construct a machine which continuously delivers work without the
expenditure of energy of some another form simultaneously.
• Heat and work are mutually convertible, but since energy can neither be created nor
destroyed, the total energy associated with an energy conversion remains constant.
The first law of thermodynamics has no theoretical proof, but experimental evidences have
confirmed its validity.
The law has not yet been violated and therefore, it is accepted as a law of nature.
This law is applicable to reversible as well as irreversible processes since no restriction was
imposed which limited its application to reversible energy transformations.
The first law of thermodynamics is concerned with two modes of energy transfer, namely,
heat and work.
(2.1)
Chemical Engineering Thermodynamics - I 2.2 First Law of Thermodynamics

FIRST LAW OF THERMODYNAMICS FOR A CLOSED SYSTEM


UNDERGOING A CYCLE / CYCLIC PROCESS
A process or series of processes, undergone by a system as a result of which the system
returns exactly to its original state, is called a cycle or cyclic process.
The first law of thermodynamics states that if a system undergoes a cycle transferring heat
and work across its boundary, then net work transfer is equal to net heat transfer. That is,

O dQ = ⌠
⌡ ⌡
O dW … (2.1)

where ⌠

O denotes the integral (or algebraic summation) over the entire cycle.

The first law of thermodynamics can also be stated as follows :


When a system undergoes a cycle then the net heat supplied/transferred to the system from
the surroundings is equal to the net work done by the system on the surroundings. That is,

O dQ = ⌠
⌡ ⌡
O dW
From surroundings On surroundings

This can also be written as


(Σ Q)cycle = (Σ W)cycle ... (2.2)

Q2
Boiler Turbine W1
3360 kJ/kg
844.221 kJ/kg

W2 Feed
pump Condenser Q1
4.221 kJ/kg
2520 kJ/kg

Fig. 2.1 : A thermal power plant (closed system)



O dQ = Q2 – Q1



O dW = W1 – W2


⌡ ⌠
O dQ = ⌡
O dW

Q2 – Q1 = W1 – W2

3360 – 2520 = 844.221 – 4.221


840 = 840
Chemical Engineering Thermodynamics - I 2.3 First Law of Thermodynamics

FIRST LAW OF THERMODYNAMICS FOR A CLOSED SYSTEM


UNDERGOING A CHANGE OF STATE (NON-FLOW PROCESS)
If a system undergoes a change of state (a process) during which energy is transferred both as
heat and work across its boundary, the net energy transfer will be stored within the system.
If Q is the amount of heat transferred to the system, i.e., the amount of heat absorbed or taken-up
by the system from its surroundings and W is the amount of work done by the system on the
surroundings during the process, then the net energy transfer to the system, i.e., the net gain of
energy by the system is Q – W.
When work is done by the system, it loses energy as it performs work at the expense of its
energy and when heat is taken-up by the system it gains energy. In our case, since the system
loses energy W, because of work done by it and gains energy Q, by transfer of heat to it from the
surroundings, the net gain of energy is Q – W. According to the first law of thermodynamics, this
net gain of energy, Q – W must be equal to the increase in the total energy of the system or the
change in the total energy of the system.
Therefore, ∆E = Q – W ... (2.3)
where ∆E - Increase in the total energy of the system
or Q = ∆E + W ... (2.4)
Q, W and ∆E are all expressed in the same units [e.g., in joules (J) in the SI system].
Boundary

W
System

Q
Surroundings

Fig. 2.2 : Heat and work interactions of a system with


its surroundings in a process (Non-flow process)
Equation (2.3) may be rewritten in differential form as
dE = dQ – dW ... (2.5)
If there are many heat and work quantities involved in the process, i.e., if there are more than
two energy interactions are involved in the process, as shown in Fig. 2.3, the first law of
thermodynamics gives

∆E = Σ Q – Σ W
⇒ ∆E = (Q1 + Q3 – Q2) – (W2 – W1 – W3)

(Q1 + Q3 – Q2) = ∆E + (W2 – W1 – W3)


Chemical Engineering Thermodynamics - I 2.4 First Law of Thermodynamics

W2

Q3 Q2
W
System
Boundary

Q1 W3

W1
Surroundings

Fig. 2.3 : System-surroundings interaction in a process


involving many heat and work quantities
The total energy or simply energy (E) of a system, neglecting magnetic, electrical and surface
energies, is given by
E = PE + KE + U ... (2.6)
where PE, KE and U refer to the potential, kinetic and internal energy respectively.
The change in the total energy is given by
∆E = ∆ PE + ∆ KE + ∆U ... (2.7)
Combining Equations (2.3) and (2.7), we get
∆ PE + ∆ KE + ∆U = Q – W ... (2.8)
Closed systems undergo non-flow processes. In case of a steady-state non-flow process, there
are no changes in the potential and kinetic energies of the system. Therefore, for such a process,
Equation (2.8) becomes
∆U = Q – W ... (2.9)
Equation (2.9) is applicable to non-flow processes involving finite changes in the system.
For differential changes, the above equation is written as
dU = dQ – dW ... (2.10)
Equations (2.9) and (2.10) are the mathematical statements of the first law of
thermodynamics for non-flow processes.
Equations (2.9) and (2.10) are applicable to reversible as well as irreversible changes,
provided that the change in the total energy of a system equals the change in the internal energy
of the system.
In Equation (2.10), dW is
dW = dWdisplacement + dWshaft + dWelectrical + ......
considering the different forms of work which may be present.
Chemical Engineering Thermodynamics - I 2.5 First Law of Thermodynamics

If only displacement work / work of expansion is present, then dW is equal to P dV and


Equation (2.10) thus becomes
dU = dQ – P dV ... for reversible process only ... (2.11)
Or, in the integral form,

∆U = Q – ⌠
⌡ P dV ... (2.12)

Note that ⌠
⌡ dW = W, ⌠
⌡ dQ = Q, but ⌠
⌡ dU = ∆U = U2 – U1
Also the difference Q – W or dQ – dW = ∆U.
Since W and Q are path functions, whereas U is a point or state function, the differentials of
W and Q are inexact differentials and thus the integration of dQ or dW results in a finite quantity
Q or W, whereas the differential of U is an exact differential and thus the integration of dU
results in a finite change between two values of U, i.e., ∆U = U2 – U1. The difference of two
path functions Q – W or dQ – dW does not depend on the path followed as it is determined
only by the initial and final states of the system and is thus equal to the difference in a state
function, U.
The first law of thermodynamics states that the change in the total energy of a system is
equal to the net energy transfer across the boundary of the system when the system undergoes a
change of state or a process.
ENERGY - A PROPERTY OF THE SYSTEM
We will now prove that the energy (E) of a system is a property of the system.
Consider a system which undergoes a change of state from state 1 to state 2 by following the
path A as shown in Fig. 2.4. The system is then restored to its exact original state-state 1 from
state 2 by following the path B. Therefore, the system undergoes a cycle. (In states 1 and 2, the
thermodynamic co-ordinates are the observable properties such as P and V that determine E.)

P C
B
A

V
Fig. 2.4 : Energy - a property of a system
For path A, according to the first law of thermodynamics, we can write
QA = ∆EA + WA
∴ ∆EA = QA – WA ... (2.13)
Chemical Engineering Thermodynamics - I 2.6 First Law of Thermodynamics

For path E, we can write


QB = ∆EB + WB
∴ ∆EB = QB – WB ... (2.14)
The path / processes A and B together constitute a cycle as by a combination of A and B the
system is restored to its initial state 1. For this cycle, we have

O dW = ⌠
⌡ ⌡
O dQ

or (Σ W)cycle = (Σ Q)cycle
∴ WA + WB = QA + AB
∴ QA – WA = – (QB – WB) ... (2.15)
Combining Equations (2.13), (2.14) and (2.15) yields
∆EA = – ∆EB ... (2.16)
The total energy change of the system along path A is identical in magnitude, but opposite in
sign, to that along the path B.
If the system returns to its exact original state - state 1 from state 2 by following the path C
instead of path E, then for this new cycle comprising paths A and C, we have
WA + WC = QA + QC
QA – WA = – (QC – WC)
∆EA = – ∆EC ... (2.17)
The total energy change of the system along path C is identical - equal in magnitude, but
opposite in sign to that along path A.
From Equations (2.16) and (2.17), it follows that
∆EB = ∆EC
If the system returns to its exact original state - state 1 from state 2 by following the path D
then on the similar line for this cycle comprising paths A and D, we can write
∆EA = – ∆ED ... (2.18)
From Equations (2.16), (2.17) and (2.18), it follows that
∆EB = ∆EC = ∆ED ... (2.19)
The paths B, C and D are different routes for changing the state of the system from state 2 to
state 1. The total energy changes in the different paths B, C and D between the two given states 1
and 2 are all equal to each other [Equation (2.19)]. The change in energy of a system between the
two given states of the system is the same, whatever path the system may follow in undergoing
that change of state. Therefore, the change in energy of a system undergoing a change of state
depends only on the initial and final states and is independent of the path followed. If some
arbitrary value is assigned to E in state 2 then the value of E in state 1 is fixed, independent of
the path followed by the system. Thus, energy has a definite value for every state of the system.
Hence, the energy of a system is a point function / state function and a property of the system.
Chemical Engineering Thermodynamics - I 2.7 First Law of Thermodynamics

ENERGY OF AN ISOLATED SYSTEM


An isolated system is one which can exchange neither mass nor energy with its surroundings.
So this system does not exchange energy as heat as well as work with its surroundings.
Therefore, for an isolated system,
dQ = 0 and dW = 0
Therefore, the first law of thermodynamics gives
dE = dQ – dW ... Equation (2.5)
dE = 0 – 0 = 0, dE = 0 ⇒ E2 = E1
or E = Constant
That is, the energy of an isolated system always remains constant or energy of an isolated
system is always conserved. The universe can be treated as an isolated system and so, we can say
that the energy of the universe is conserved.
PERPETUAL MOTION MACHINE OF THE FIRST KIND - PMM1
The first law of thermodynamics for a system undergoing a cyclic change is

O dQ = ⌠
⌡ ⌡
O dW ... Equation (2.1)

That is, work done by a system during a cyclic process is equal to the heat transferred to the
system from its surroundings. In a cyclic process, if there is no transfer of heat to a system, the
system does not perform work.
i.e.‚ if ⌠ O dW = 0
O dQ = 0‚ then according to Equation (2.1)‚ ⌠
⌡ ⌡
 
A cyclically operating machine which continuously delivers work without corresponding
expenditure of energy simultaneously is called a perpetual motion machine of the first kind
(PMM1).

We know that if ⌡ ⌠
O dQ = 0, then ⌡
O dW = 0. Therefore, a PMM1 is impossible. Such a
machine violates the first law of thermodynamics. A PMM1 creates energy in the form of work
out of nothing. It is thus a fictitious machine.
Q

Engine W

Fig. 2.5 : A PMM1


Chemical Engineering Thermodynamics - I 2.8 First Law of Thermodynamics

CYCLIC PROCESS
A system undergoing a cyclic process, after completion of the process, returns to its exact
original state. Therefore, the total energy of the system will remain unchanged so that ∆E must
be zero. Hence, it follows from Equation (2.3) that the work done by the system is equal to the
heat absorbed by the system in the process.
i.e., Q = W ... (2.20)
This equation is an expression indicating that a PMM1 is impossible. Provided the system
does not undergo any net change, i.e., ∆E is zero, it is impossible to produce work without
extracting energy, namely heat from an outside source. In any cycle or cyclic process, ∆E is zero
and hence the heat absorbed by the system is equal to the work done by the system.
ISOTHERMAL PROCESS
In this process, temperature remains constant.
Since temperature is a measure of the internal energy, the change in the internal energy in
this process is zero.
According to the first law,
∆U = Q – W
0 = Q–W
∴ Q = W ... (2.21)
That is, heat absorbed by the system gets converted to work.
ADIABATIC PROCESS
In an adiabatic process, Q = 0 ... no heat interaction with the surroundings. Therefore,
according to the first law of thermodynamics,
∆U = Q – W
∆U = 0 – W
∴ W = – ∆U ... (2.22)
Hence, in an adiabatic process, the work is done at the expense of the internal energy, so the
internal energy decreases.
CONSTANT VOLUME PROCESS
In this process, the volume remains constant. Therefore, dV = 0.
According to the first law of thermodynamics,
∆U = Q – P dV
∴ ∆U = Q – 0
∴ ∆U = Q ... (2.23)
Therefore, the amount of heat supplied or absorbed by the system will be used to increase the
internal energy.
Chemical Engineering Thermodynamics - I 2.9 First Law of Thermodynamics

ENTHALPY OF A SYSTEM
The heat supplied to a system at constant volume is equal to the change in internal energy
(dU = dQ – dW, dW = P dV, dV = 0 ∴ dU = dQ - in the integral form ⇒ ∆U = Q). When
the volume of the system changes against a constant external pressure, the change in U is not
equal to the heat supplied since a part of this energy is used by the system to occupy a new
volume. However, the heat supplied at constant pressure can be measured as the change in
another thermodynamic property of the system to which is given the name enthalpy and the
symbol H. It is the most widely used thermodynamic function.
The enthalpy of a system is defined mathematically as
H = U + PV ... (2.24)
It is the sum of internal energy and pressure - volume product. It is the total energy of the
system.
The product PV as U has the units of energy. Therefore, H also has the units of energy.
As P, V and U all are state functions, H is also a state function / point function. The enthalpy H
of a system is an extensive property of the system.
In the differential form, Equation (2.24) can be written as
dH = dU + d (PV) ... (2.25)
In the integral form, this equation becomes
∆H = ∆U + ∆ (PV) ... (2.26)
This is applicable to any finite change taking place in the system. The change in ∆H depends
only on the initial and final states of the system.
Now, we will obtain a relation between ∆H and Q (heat supplied to a system) for a
mechanically reversible, non-flow process.
For a reversible process, we have
dW = P dV ... (2.27)
The first law of thermodynamics for a non-flow process, in the differential form, is
dU = dQ – dW ... (2.28)
Combining Equations (2.27) and (2.28) gives
dU = dQ – P dV ... (2.29)
We have, H = U + PV
In the differential form it becomes
dH = dU + d (PV)
This equation can be expanded as
dH = dU + P dV + V dP
For a constant pressure process, dP = 0 ∴ V dP = 0 and the above equation becomes
dH = dU + P dV ... (2.30)
Chemical Engineering Thermodynamics - I 2.10 First Law of Thermodynamics

Substituting for dU from Equation (2.29) into Equation (2.30), we get


dH = dQ – P dV + P dV
dH = dQ ... at constant pressure ... (2.31)
or in the integral form,
∆H = Q ... at constant pressure ... (2.32)
Thus, the change in enthalpy of a system for a reversible, non-flow process at constant
pressure is equal to the heat supplied / added to the system.
For a constant volume non-flow process, we have
P dV = 0 ... the work of expansion is zero
∴ dU = dQ ... at constant volume ... (2.33)
In the integral form, it becomes
∆U = Q ... (2.34)
The change in internal energy of a system is equal to the heat supplied / added to the system
for a reversible non-flow process occurring at constant volume.
HEAT CAPACITY
The quantity of heat required to raise the temperature of 1 kg of substance by 1 K (one
degree kelvin) is defined as heat capacity. Usually, the heat capacity is defined on a unit mass
basis or on a unit mole basis. When the heat capacity is defined on a unit mole basis, i.e., for one
mole of a pure substance, it is known as molar heat capacity. The molar heat capacity is denoted
by the letter C and has the units of kJ/(kmol·K).
Since the heat capacity varies with temperature, the true molar heat capacity at any
temperature T is given by the differential equation
dQ
C = dT ... for 1 mole ... (2.35)
Since Q depends on the path taken, the heat capacity will also be uncertain, unless conditions
are specified, such as constant volume or constant pressure, which define the path. Therefore,
we have heat capacities at constant pressure and at constant volume.
MOLAR HEAT CAPACITY AT CONSTANT PRESSURE
At constant pressure, Equation (2.35) may be written as
dQ
Cp = dT ... (2.36)
where Cp is the molar heat capacity at constant pressure.
∴ dQ = Cp dT ... (2.37)
For a reversible non-flow process, we have
dU = dQ – dW
dU = dQ – P dV
dQ = dU + P dV
Chemical Engineering Thermodynamics - I 2.11 First Law of Thermodynamics

For a constant pressure process, P dV = d (PV) . Therefore, the above equation becomes
dQ = dU + d (PV) = dH ... (2.38)
Combining Equations (2.37) and (2.38) gives
dH = Cp dT
dH
Cp = dT ... at constant P

or
∂H
Cp =   ... (2.39)
 ∂T P
Since H, T and P are properties, Cp is a property of the system.
The partial differential notation is used because H is a function of the pressure as well as the
temperature. The subscript P as applied to the partial derivative in Equation (2.39) indicates that
in this case the pressure is kept constant.
It is clear from Equation (2.39) that the (molar) heat capacity of a system at constant pressure
is equal to the rate of increase of enthalpy with temperature at constant pressure.
For a constant pressure process, we have
dQ = dH = Cp dT ... (2.40)
At constant pressure, the enthalpy change of a system is thus given by
T2

Q = ∆ H = H2 – H1 = ⌠
⌡ Cp dT ... (2.41)
T1
If Cp is constant, then Equation (2.41) gives
∆H = Cp (T2 – T1) ... (2.42)
MOLAR HEAT CAPACITY AT CONSTANT VOLUME
At constant volume, Equation (2.35) may be written as
dQ
Cv = dT ... (2.43)
∴ dQ = Cv dT ... (2.44)
where Cv is the molar heat capacity at constant volume.
We have dU = dQ – dW ... for a non-flow process
For a reversible process,
dW = P dV
∴ dU = dQ – P dV
For a constant volume process, P dV = 0 as dV = 0.
∴ dU = dQ ... for a constant volume process ... (2.45)
Combining Equations (2.44) and (2.45) gives
dU = Cv dT
dU
i.e., Cv = dT ... at constant V

Cv =  
∂U
or ... (2.46)
∂TV
Therefore, the (molar) heat capacity of a system at constant volume is equal to the rate of
increase of internal energy with temperature at constant volume.
Chemical Engineering Thermodynamics - I 2.12 First Law of Thermodynamics

For a constant volume process,


dQ = dU = Cv dT
In the integral form, it becomes
T2

Q = ∆U = ⌠
⌡ Cv dT ... (2.47)
T1

If Cv is constant, then the above equation becomes


Q = ∆U = Cv (T2 – T1) ... (2.48)
We have, Q = ∆H ... for constant pressure process
and Q = ∆U ... for constant volume process
Therefore, for a reversible, constant pressure, non-flow process, the heat transferred is equal
to the enthalpy change of the system; while for a reversible, constant volume, non-flow process,
the heat transferred is equal to the internal energy change of the system.
REVERSIBLE CONSTANT VOLUME PROCESS
In a constant volume process, the working substance is contained in a rigid vessel, therefore,
the boundaries of the system are immovable and consequently, there is no work of expansion.
The work done in a reversible, non-flow process is given by

W = ⌠
⌡ P dV
For a constant volume process, as dV = 0, the work done by the gas during this process is
zero.

W = ⌠
⌡ P dV = 0
Consider a certain amount of an ideal gas contained in a rigid vessel. The vessel is kept in
thermal contact with a hot body so that the heat is supplied to the gas. Due to the addition of
heat, the temperature as well as pressure of the gas increases since its volume is constant. This
process is called a constant volume or isochoric or isometric process. As a result of heat addition,
gas changes its state from state 1 (P1, T1) to state 2 (P2, T2). The path 1-2 followed by the gas can
be shown on a P-V diagram as shown in Fig. 2.6.

T2

P 1

T1

V
Fig. 2.6 : P-V diagram of a constant volume process
Chemical Engineering Thermodynamics - I 2.13 First Law of Thermodynamics

For a non-flow process (closed systems), the first law of thermodynamics is


∆U = Q – W
Since W = 0 for a constant volume process, it reduces to
∆U = Q
or U2 – U1 = Q ... (2.49)
Therefore, in a constant volume process, the heat added to a system is equal to the increase in
the internal energy of the system.
For this process, the heat added is equal to the product of molar heat capacity and rise in
temperature. (dQ = Cv dT). Therefore,
T2

∆U = U2 – U1 = Q = ⌠
⌡ Cv dT ... for 1 mole ... (2.50)
T1

Equation (2.50) applies to all gases, whether ideal or real.


For an ideal gas, for any process, the change in internal energy is given by
T2

∆U = ⌠
⌡ Cv dT ... (2.51)
T1

In the differential form, the above equations are written as


dW = P dV = 0 since dV = 0
dU = dQ – dW
dU = dQ
dU = Cv dT

REVERSIBLE CONSTANT PRESSURE PROCESS


For a constant pressure process, the boundary of the system moves against a constant
external pressure as heat is supplied. Consider a certain amount of an ideal gas enclosed in a
frictionless piston - cylinder assembly (which moves without friction). The piston is loaded with
a certain mass, which exerts a constant pressure on the gas. Initially assume that the pressure
exerted by the mass loaded is exactly equal to the gas pressure. So the piston will not move
upward or downward as the pressure forces across the piston are in exact balance. The assembly
is now placed in contact with a hot body. Due to transfer of heat from the body, the gas expands
and performs work on the surroundings. The temperature of the gas also increases. If the transfer
of heat occurs with no temperature difference between the system and its surroundings, the
process 1-2 followed by the system is reversible. This process can be represented on a P-V
diagram as shown in Fig. 2.7.
Chemical Engineering Thermodynamics - I 2.14 First Law of Thermodynamics

P 1 T2
2

T1

V
Fig. 2.7 : P-V diagram for a constant pressure process
(1 to 2 - constant pressure heating of a gas)
For a constant pressure process (non-flow), the work done by 1 mole of the gas is given by
V2 V2

W = ⌠
⌡ P dV = P ⌠
⌡ dV = P (V2 – V1) ... for 1 mole ... (2.52)
V1 V1

For a non-flow process, we have


dU = dQ – dW
dQ = dU + dW = dU + P dV
For a constant pressure, it becomes
dQ = dU + d (PV) ... since V dP = 0 at constant P
dQ = d (U + PV)
dQ = dH ... since H = U + PV, differential form
or Q = ∆H ... integrated form ... (2.53)
Thus, in a constant pressure process, the heat added is equal to the change in the enthalpy of
the system.
When a system undergoes a constant pressure process, there will be a change in the
temperature of the system. The heat supplied to the system is equal to the product of molar heat
capacity at constant pressure and rise in temperature.
∴ dQ = Cp dT
∴ dQ = dH = Cp dT ... (2.54)
T2

Integrating, we get Q = ∆H = ⌠
⌡ Cp dT ... for 1 mole ... (2.55)
T1

If Cp is constant, then the above equation gives


Q = ∆H = Cp (T2 – T1) ... (2.56)
Chemical Engineering Thermodynamics - I 2.15 First Law of Thermodynamics

The work done by the system during a constant pressure process is given by
W = P (V2 – V1)
W = P2 V2 – P1 V1 ... since P2 = P1 = P
W = RT2 – RT1 ... since PV = RT for an ideal gas
W = R (T2 – T1) ... for 1 mole of an ideal gas ... (2.57)
REVERSIBLE CONSTANT TEMPERATURE (ISOTHERMAL) PROCESS
A process in which the temperature of a system is kept constant is called an isothermal
process.
For an ideal gas, U = f (T) only, i.e., U does not change unless T changes. Therefore, for an
isothermal process (T = constant) involving an ideal gas, the change in internal energy of the
gas is zero.
∆U = 0 ... for an ideal gas undergoing isothermal process
The first law equation for a non-flow process is
dU = dQ – dW
or ∆U = Q – W
For an isothermal process involving an ideal gas (for which U is a function of T only), the
above equation becomes
0 = Q–W
∴ Q = W
The work done by a system in a reversible non-flow process is given by
V2

W = ⌠
⌡ P dV
V1

∴ The work done by 1 mole of an ideal gas in a reversible, isothermal non-flow process is
given by
V2

W = ⌠
⌡ P dV
V1

The ideal gas equation is


PV = RT ... for 1 mole
RT
∴ P = V
V2 V2
RT dV
∴ W = ⌠
⌡ V dV = RT ⌠
⌡ V
V1 V1

Since R and T are constant, they are taken outside of the integral sign.
V2
W = RT ln V ... for 1 mole ... (2.58)
1
Chemical Engineering Thermodynamics - I 2.16 First Law of Thermodynamics

For an ideal gas, PV = RT


At constant T, P1V1 = RT = P2 V2
V2 P1
∴ V1 = P2 ... (2.59)

From Equations (2.58) and (2.59), we get


P1
W = RT ln P ... for 1 mole ... (2.60)
2

Since Q = W ... for an isothermal process involving an ideal gas, it follows that
V2 P1
Q = RT ln V = RT ln P ... for 1 mole ... (2.61)
1 2

REVERSIBLE ADIABATIC PROCESS (EXPANSION OR COMPRESSION)


A process in which no heat enters or leaves the system is called an adiabatic process.
In this process, there is no heat transfer / heat interaction between the system and its
surroundings, i.e., dQ = 0 or Q = 0.
Consider a certain amount of gas contained in an insulated piston - cylinder assembly at high
pressure (since the assembly is insulated, there is no heat transfer / interaction).
If the external pressure is gradually reduced, the gas in the cylinder expands slowly in such a
way that the pressure forces across the piston are always in exact balance. Thus, the gas
undergoes a reversible adiabatic process (expansion).
A reversible, adiabatic process is known as an isentropic process (a process with constant
entropy). A reversible process is isentropic when Q = 0 (∆S = Q/T, S = entropy, when Q = 0,
∆S = 0), that is, a reversible adiabatic process is isentropic.
The application of the first law of thermodynamics to 1 mole of a gas undergoing a non-flow,
reversible process gives
dU = dQ – dW = dQ – P dV
This equation for a reversible, adiabatic process (for which dQ = 0) becomes
dU = – dW = – P dV
– dU = dW = P dV
or – ∆U = W = ⌠
⌡ P dV
Thus, the work done by a system during an adiabatic process (reversible adiabatic expansion)
is equal to the decrease in the internal energy of the system. In this process, as no heat enters or
leaves a system, the work is done by the system at the expense of the internal energy so the
internal energy of the system (the gas) decreases and consequently the temperature of the system
falls.
For an ideal gas, U = f (T) only or dU = Cv dT
The equation dU = Cv dT holds good for an ideal gas for any process.
Chemical Engineering Thermodynamics - I 2.17 First Law of Thermodynamics

For an ideal gas in a reversible, adiabatic process, we have


– dU = dW = P dV
– dU = P dV
– Cv dT = P dV ... (2.62)
From Equation (2.62), it is clear that in a reversible, adiabatic process, the signs associated
with dV and dT are opposite. Thus, if the volume of the gas is increased, as in an adiabatic
expansion process, the temperature of the gas must fall (and the reverse of this statement is true
for an adiabatic compression - in a process of compression the temperature of the gas must rise).
This is expected since a gas does work when it expands, its internal energy must decrease, as
during an adiabatic process no heat enters or leaves the system. As the work is done at the
expense of its internal energy, the decrease of energy - expansion of the gas is accompanied by a
fall of temperature. (During an adiabatic expansion the temperature of the gas decreases. During
a compression process, the work is done on the system, it is stored in the system in the form of
internal energy and consequently the temperature of the system increases.)
Suppose 1 mole of an ideal gas undergoes a reversible, adiabatic process from the initial state
(P1, V1, T1) to the final state (P2, V2, T2).
We have, – dU = dW = P dV
– Cv dT = P dV = dW
or W = – Cv (T2 – T1) ... (2.63)
For 1 mole of an ideal gas undergoing a reversible adiabatic process, the first law gives
– Cv dT = P dV ... (2.64)
RT
For an ideal gas, PV = RT orP = V

Equation (2.64) on substitution for P = V  becomes


RT
 
RT
– Cv dT = V dV
dT R dV
Rearranging gives T = – Cv · V ... (2.65)
The ratio of the molar heat capacities at constant pressure and constant volume, Cp/Cv is
designated by the symbol γ (adiabatic index).
Cp
∴ Cv = γ
or Cp = Cv · γ ... (2.66)
For an ideal gas, the relationship between Cp and Cv is given by
Cp – Cv = R ... (2.67)
where R is the universal gas constant [8.31451 J/(mol·K)].
Chemical Engineering Thermodynamics - I 2.18 First Law of Thermodynamics

Combining Equations (2.66) and (2.67) gives


Cv γ – Cv = R
Cv = Cv · γ – R
Cv (γ – 1) = R
R
∴ Cv = ... (2.68)
γ–1
Substituting for Cv from Equation (2.68), Equation (2.65) becomes
dT –R dV
T = R/(γ – 1) · V
dT dV
T = – (γ – 1) V ... (2.69)

Integrating Equation (2.69) between the initial (P1, V1, T1) and the final (P2, V2, T2)
conditions gives
T2 V2

⌠ dT dV
⌡ T = – (γ – 1) ⌠
⌡ V
T1 V1

T2 V2
ln T = – (γ – 1) ln V
1 1
T2 V1
∴ ln T = (γ – 1) ln V ... (2.70)
1 2

Rearranging Equation (2.70) gives us the relationship between temperature and volume in an
adiabatic process.
T2 V1γ–1
= V 
T1  2
or TVγ–1 = Constant = C ... (2.71)
This equation can be utilized to determine the temperature change in a reversible adiabatic
process for an ideal gas. It relates temperature and volume for an ideal gas with constant heat
capacities undergoing a reversible adiabatic process.
For 1 mole of an ideal gas undergoing a reversible adiabatic process for the initial state
(P1, V1, T1) to the final state (P2, V2, T2) derive a relationship between P and V.
For 1 mole of an ideal gas undergoing a reversible adiabatic process, the first law gives
– Cv dT = P dV ... (2.72)
RT
For an ideal gas, PV = RT or P = V
RT R
Substituting V for P and for Cv, Equation (2.72) becomes
(γ – 1)
R RT
– dT = V dV
(γ – 1)
Chemical Engineering Thermodynamics - I 2.19 First Law of Thermodynamics

Rearranging gives
dT dV
T = – (γ – 1) V ... (2.73)

The ideal gas law is


PV = RT
∴ P dV + V dP = R dT
P dV + V dP
dT = R ... (2.74)

PV
and T = R ... (2.75)

Substituting for dT from Equation (2.74) and T from Equation (2.75), Equation (2.73)
becomes
P dV + V dP dV
R (PV/R) = – (γ – 1) V
dV dP dV
∴ V + P = – (γ – 1) V
dP dV dV
P = – (γ – 1) V – V
dP dV dV
P = (– γ + 1 – 1) V = – γ V ... (2.76)

Integrating Equation (2.76) between the limits P1, V1 and P2, V2, the initial and final
pressures and volumes, respectively, we get
P2 V2

⌠ dP dV
⌡ P = –γ⌠
⌡ V
P1 V1

P2 V2
ln P = – γ ln V
1 1

P2 V1
or ln P = γ ln V
1 2

P2 V1γ
i.e., P1 = V2
γ
or PV = Constant ... (2.77)
This equation relates pressure and volume for a reversible adiabatic process involving an
ideal gas.
Equation (2.77) is the law of reversible, adiabatic process.
A reversible, adiabatic process followed by an ideal gas can be represented on a P-V diagram
as shown in Fig. 2.8.
Chemical Engineering Thermodynamics - I 2.20 First Law of Thermodynamics

P
2

V
Fig. 2.8 : Reversible adiabatic path on P-V diagram
T2 V1γ–1
We have, T1 = V2 ... (2.78)
and PV = RT
∴ P1V1 = RT1 and P2 V2 = RT2
V1 RT1 P2 T1 P2
∴ V2 = P1 RT2 = T2 × P1
V1
Substituting for V , Equation (2.78) becomes
2
T2 P2 × T1γ–1 = P2γ–1 T1γ–1
= P T  P  T 
T1  1 2  1  2
T2 1 P2γ–1
× = P 
T1 T1 γ –1  1
T 
 2
T21+ (γ–1) = P2γ–1
∴ T  P 
 1  1
T2γ = P2γ–1
∴ T  P 
 1  1
This can be written as
γ–1
T2 = P2 γ
T  P 
 1  1
T
i.e., (γγ–1) /γγ = Constant
P
(1–γ) /γγ
or TP = Constant ... (2.79)
dP dV
We have, P = – γ V ... Equation (2.76)

This can be rewritten as


Chemical Engineering Thermodynamics - I 2.21 First Law of Thermodynamics

d ln P = – γ d ln V
Thus, when an ideal gas undergoes a reversible adiabatic process, a plot of ln P v/s ln V
yields a straight line having a slope of – γ.

ln P Slope = -g

ln V
Fig. 2.9 : Reversible adiabatic process path on ln P vs ln V diagram
γ
or PV = C
ln P + γ ln V = ln C
∴ ln P = – γ ln V + ln C ⇒ y = mx + C with m = – γ
The work done by 1 mole of an ideal gas during a reversible, adiabatic process is given by
the expression
W = – Cv (T2 – T1) = Cv (T1 – T2)

where T1 and T2 being the initial and final temperatures.


Cv is given by

Cv = R/(γ – 1)

R
∴ W = (T1 – T2) ... (2.80)
γ–1

For an ideal gas, PV = RT


P1V1 P2V2
∴ T1 = R and T2 = R

Equation (2.80) thus becomes

R P1V1 P2V2
W = – R
γ–1 R 
P1V1 – P2V2
W = ... (2.81)
γ–1
Chemical Engineering Thermodynamics - I 2.22 First Law of Thermodynamics

Usually, V2 is not known. It can be eliminated from Equation (2.81) by making use of the
γ
law of reversible adiabatic process : PV = Constant.
γ
PV = C
γ γ
P1V1 = P2V2

P1Vγ11/γ 1/γ
V2 =  P  =
P1 · V
∴ P  ... (2.82)
 2   2 1

Substituting for V2 from Equation (2.82), Equation (2.81) becomes


1/γ
P1V1 – P2 (P1/P2) V1
W =
γ–1
Taking P1V1 common from both the terms of the numerator on the R.H.S. gives
P1V1  P2P1 
1/γ
W = 1 – ·
P 
 2 
γ – 1  P1

P1V1  
1 – (1/γ)
(P2) 
W = 1–
γ–1  1 – (1/γ)
 (P1) 
1 1 1 – (1/γ) (γ–1) /γ
P2 × 1/γ
= (P2) = (P2)
P2
 (γ–1)

P1V1  P2 γ 
1– 
γ –1  P1 
W = ... (2.83)

We know that P1V1 = RT1


With this, Equation (2.83) becomes
 (γ–1)

RT1  P2 γ 
W =  1 – P   ... (2.84)
γ–1   1 
The work done in a reversible adiabatic process is given by the shaded area under the curve
on a P-V diagram between the two ordinates at the initial and final volumes. [Fig. 2.8]. This area
can be evaluated by Equation (2.84).
V2

W = ⌠
⌡ P dV ... Work done in a reversible non-flow process
V1

γ
The law of reversible adiabatic process is PV = Constant
γ C
PV = C ⇒ P = γ
V
Chemical Engineering Thermodynamics - I 2.23 First Law of Thermodynamics

∴ The work done in a reversible adiabatic process is


V2
C
W = ⌠
⌡ γ dV
V1 V

V2 V2
dV –γ
W = C⌠
⌡ γ = C⌠
⌡ V dV
V1 V V1

V2
C –γ+1
=
–γ + 1
V [ ]
V1

C –γ+1 –γ+1
=
–γ + 1 [
V2 – V1 ]
C –γ+1 –γ+1
=
γ–1 [
V1 – V2 ] [as – γ + 1 = – (γ – 1)]
1 –γ+1 –γ+1
W =
γ–1 [
CV1 – CV2 ] ... (2.85)

γ
PV = C
γ γ
∴ C = P1V1 and C = P2V2

Thus, substituting for C, Equation (2.85) becomes


1 γ –γ+1 γ –γ+1
W = [
P V · V1 – P2V2 · V2
γ–1 1 1 ]
P1V1 – P2V2
W = ... (2.86)
γ–1
We have, P1V1 = RT1 and P2V2 = RT2 ... for an ideal gas.
Therefore, Equation (2.86) becomes
R
W = (T1 – T2) ... (2.87)
γ–1
POLYTROPIC PROCESS
This a general process without specific conditions other than mechanical reversibility.
All other processes are special cases of this general process. The law of this process is given by
PVn = Constant, PVn = C ... (2.88)
where C and n are constants, and n is called the polytropic index.
The process which follows the path described by Equation (2.88) is called the polytropic
process.
Chemical Engineering Thermodynamics - I 2.24 First Law of Thermodynamics

Special cases of a polytropic process for an ideal gas by assigning different values to n :
In a polytropic process :
(i) When n = 0, PVo = C, i.e., P = constant, the process reduces to a constant pressure
process (an isobaric process).
(ii) When n = 1, PV1 = C, PV = constant, i.e., T = constant (since PV = RT and if
T = constant, PV = constant since R is a constant), the process reduces to a constant
temperature process - an isothermal process.
(iii) When n = ∞, PV∞ = constant or P1/∞ V = constant, i.e., V = constant, the process
reduces to a constant volume process - an isochoric process.
γ
(iv) When n = γ, PV = C, the process reduces to an adiabatic process (an isentropic
process).
This is illustrated on a P-V diagram in Fig. 2.10.
n
sio

C D
es

n=¥
pr
m
Co

B
n=
n

r
=
1

n=0 n=0
P A A'
1
n=
1 B'
n
=
r
ion

C'
ns
pa
Ex

D' n=¥

V
Fig. 2.10 : P-V diagram of various polytropic processes
State 1 to State A : Constant pressure cooling (n = 0)
State 1 to State A' : Constant pressure heating (n = 0)
State 1 to State B : Isothermal compression (n = 1)
State 1 to State B' : Isothermal expansion (n = 1)
State 1 to State C : Adiabatic compression (n = γ)
State 1 to State C' : Adiabatic expansion (n = γ)
State 1 to State D : Constant volume heating (n = ∞)
State 1 to State D' : Constant volume cooling (n = ∞)
Chemical Engineering Thermodynamics - I 2.25 First Law of Thermodynamics

The general equations which are applicable to an ideal gas in a non-flow process are
applicable to a polytropic process. Thus, for 1 mole of an ideal gas undergoing a reversible
polytropic process, we have
dU = dQ – dW or ∆U = Q – W
dW = P dV or W = ⌠
⌡ P dV
dU = Cv dT or ∆U = ⌠
⌡ Cv dT
dH = Cp dT or ∆H = ⌠
⌡ Cp dT
The work done during a reversible polytropic process by 1 mole of an ideal gas is given by
P1V1 – P2V2
W = n–1 ... (2.89)
The work done during a reversible process is given by
V2

W = ⌠
⌡ P dV ... (2.90)
V1

The law of polytropic process is given by


PVn = C
C
∴ P = n ... (2.91)
V
Substituting P from Equation (2.91) into Equation (2.90) gives
V2 V2
C dV
W = ⌠
⌡ Vn dV = C ⌠
⌡ Vn
V1 V1
V2
 V–n+1  C –n+1 –n+1
Integrating gives, W = C –n + 1 = – n + 1 V2 – V1
 V [ ]
1

C –n+1 –n+1 C –n+1 –n+1


W = n – 1 – V2 + V1[ = n – 1 V1 – V2 ] [ ]
1 –n+1 –n+1
W = n – 1 CV1 – CV2[ ] ... (2.92)

We have PVn = C
n n
∴ C = P1V1 and C = P2V2
With these C values, Equation (2.92) becomes
1 n –n+1 n –n+1
[
W = n – 1 P1V1 · V1 – P2V2 · V2 ]
P1V1 – P2V2
W = n–1 ... (2.93)
We have, P1V1 = RT1 and P2V2 = RT2
Chemical Engineering Thermodynamics - I 2.26 First Law of Thermodynamics

Replacing P1V1 and P2V2 from Equation (2.93) gives


R (T1 – T2)
W = n–1 ... (2.94)

Equation (2.93) or (2.94) gives the work done by 1 mole of an ideal gas in a reversible
polytropic process.
Let us eliminate V2 from Equation (2.93), as it is usually not known.
P1V1 – P2V2
W = n–1 ... Equation (2.93)
n n
PVn = C ⇒ P1V1 = C and P2V2 = C
n n
∴ P1V1 = P2V2
n P1 n
∴ V2 = P V1
2

V2 = P  V1
P1 1/n
∴ ... (2.95)
 2
Substituting for V2 from Equation (2.95) in Equation (2.93), we get
P1V1 – P2 (P1/P2)1/n V1
W = n–1
1–1/n 1/n
P1V1 – P2 · P1 · V1
W = n–1
(n–1)/n 1/n
P1V1 – P2 · P1 V1
= n–1
(n–1)/n 1/n
P1V1 – P2 · P1 · (P1/P1) V1
= n–1
... second term is multiplied by P1/P1
Taking P1V1 common from the terms of the numerator on the R.H.S. gives
P1V1 (n–1)/n 1/n
W = n – 1 1 – P2 [ · P1 · (1/P1) ]
P1V1  P(n –1)/n

= n–1 1 – 2 
 1–1/n 
 P1 
P1V1  P(n –1)/n

= n–1 1 – 2 
 P(n–1)/n
 1 
P1V1  P2(n–1)/n
W = n–1 1 – P   ... (2.96)
  1 
Chemical Engineering Thermodynamics - I 2.27 First Law of Thermodynamics

Equation (2.96) can be used to calculate the work done in a reversible, non-flow, polytropic
process.
We have, P1V1 = RT1. Therefore, Equation (2.96) becomes
RT1  P2(n–1)/n
W = n – 1 1 – P   ... (2.97)
  1 
The change in internal energy for 1 mole of an ideal gas during a reversible polytropic
process is given by
dU = Cv dT ... valid for an ideal gas for any process
If Cv is taken as constant, integration of the above equation yields
∆U = Cv (T2 – T1) ... (2.98)
Cp
We have, Cv = γ
∴ Cp = γ · Cv
Cp – Cv = R ... for an ideal gas
R
∴ γCv – Cv = R ⇒ Cv =
(γ – 1)
Substituting for Cv, Equation (2.98) becomes
R
∆U = (T2 – T1) ... (2.99)
γ–1
HEAT TRANSFER DURING A REVERSIBLE POLYTROPIC PROCESS
We have, ∆U = Q – W
∴ Q = ∆U + W ... (2.100)
Substituting the values of ∆U from Equation (2.99) and of W from Equation (2.94) into
Equation (2.100), we get
R R
Q = (T2 – T1) + n – 1 (T1 – T2)
γ–1
This can be rewritten as
R R
Q = n – 1 (T1 – T2) – (T – T2)
γ–1 1

= R (T1 – T2) n – 1 –
1 1 
 γ – 1
(γ – 1) – (n – 1)
= R (T1 – T2)  
 (n – 1) (γ – 1) 
R (T1 – T2) (γ – 1 – n + 1)
=
(n – 1) (γ – 1)
(γ – n) R (T1 – T2)
= × (n – 1)
(γ – 1)
Chemical Engineering Thermodynamics - I 2.28 First Law of Thermodynamics

Q = 
γ – n since W = R (T1 – T2)
or  × W,
 n–1  ... (2.101)
γ – 1 
γ – n × Work done in a reversible 
i.e., Q =  
γ – 1 non-flow polytropic process
FIRST LAW OF THERMODYNAMICS FOR STEADY-STATE FLOW PROCESS
(OPEN SYSTEM)
In the chemical process industry, we come across a variety of process equipments such as
reactors, absorption towers, distillation columns, etc. as well as machines such as pumps,
compressors, etc. through which there is a continuous flow of material in and out. Such processes
are called flow processes and such devices operate as open systems.
A process in which material continuously flows in and out of a system through its boundary
is called as a flow process.
Most of the devices we come across are operated under steady-state conditions - conditions
which are invariant with time.
A steady-state flow process is one in which the conditions at all points in the apparatus are
constant with time, i.e., invariant with time. A steady-state flow process must satisfy the
following requirements :
(i) The rate of mass flow into the apparatus is always equal to the rate of mass flow out of
the apparatus (system), so that the quantity of material within the apparatus is constant,
i.e., there is no accumulation or depletion of the material in the apparatus over the
period of the time considered.
(ii) The conditions such as temperature, pressure and composition of streams flowing at
any point within the apparatus are constant.
(iii) The rates of heat transfer and work transfer across the boundary of the system are
constant with time.
Thus, a heat exchanger operates in steady-state flow if the flow rates, compositions,
temperatures, etc. of all entering as well as leaving streams are constant.
Consider a steady-state flow system as shown in Fig. 2.11.
Section 2

Turbine Ws z2

Section 1 Heat exchanger

z1
Q Datum level

Fig. 2.11 : Steady-state flow process


Chemical Engineering Thermodynamics - I 2.29 First Law of Thermodynamics

A fluid, either liquid or gas flows through the system from section 1 to section 2.
At section 1, let U1, u1, V1, P1 and z1 be the specific internal energy, velocity, specific
volume, pressure and elevation above an arbitrary datum level of the fluid, respectively.
[U1 in J/kg, u1 in m/s, V1 in m3/kg, P1 in N/m2 and z1 in m.]
Similarly, at section 2, let U2, u2, V2, P2 and z2 be the specific internal energy, velocity,
specific volume, pressure and elevation of the fluid, respectively.
·
Let m be the mass flow rate of the fluid (kg/s) which is constant from entrance to exit.
In our case, the only heat interaction is the heat added to the flow system in the heat
exchanger.
·
Let Q be the rate of heat transfer to the flow system from the surroundings (J/s).
In our case the only work interaction with surroundings is the work done / work performed
by the turbine (shaft work).
·
Let W s be the rate of work done by the flow system in J/s. In a flow process, the work
involved is of two types : the external work and the flow work. In our case, the external work
occurs in the form of shaft work (Ws).
Flow energy : It is also called as pressure energy or flow work and comes into the picture
when we are dealing with flow processes. It represents the work necessary to advance / propel a
fluid against the existing pressure. A source of the flow energy is a pump, blower or compressor
which imparts energy to the fluid.
The flow energy is the product of the pressure and specific volume of the fluid. Therefore,
the flow energy with unit mass of fluid is PV.
Flow energy at section 1 = P1V1 , J/kg
Flow energy at section 2 = P2V2
At section 1 (i.e., at entrance),
Internal energy with = U , J/kg
 unit mass of fluid  1

1 2
Kinetic energy = 2 u1
Potential energy = gz1
Flow energy = P1V2
The total energy at section 1 (i.e., at the entrance) with unit mass of fluid entering into the
flow system is
1 2
U1 + 2 u1 + gz1 + P1V1, J/kg
The rate of flow of total energy into the flow system is
m U1 + 2 u1 + gz1 + P1V1
· 1 2
 
Chemical Engineering Thermodynamics - I 2.30 First Law of Thermodynamics

At section 2 :
Internal energy with = U , J/kg
 unit mass of fluid  2

1 2
Kinetic energy = 2 u2

Potential energy = gz2


Flow energy = P2V2
The total energy at section 2 (i.e., at the exit) with unit mass of fluid leaving the flow
system is
1 2
U2 + 2 u2 + gz2 + P2V2 , J/kg
The rate of flow of total energy out of the flow system is
m U2 + 2 u2 + gz2 + P2V2
· 1 2
 
For this steady-state flow process, the first law of thermodynamics is
 Rate of total   Rate of total 
 +  inflow as heat  = energy outflow + outflow as work
Rate of energy Rate of energy
energy inflow ... (2.102)
 at entrance   at exit 
Substituting the values of the terms involved, Equation (2.102) becomes
· ·
m U1 + 2 u1 + gz1 + P1V1 + Q = m U2 + 2 u2 + gz2 + P2V2 + Ws
· 1 2 · 1 2
   
· ·
U + 1 u21 + gz + P V  + Q = U + 1 u22 + gz + P V  + Ws ... (2.103)
 1 2 1 1 1
 ·  2 2 2 2 2
 ·
m m
·
Ws J/s
But Ws = · = Shaft work (energy transfer as work) , kg/s = J/kg
m
·
Q J/s
and Q = · = Heat flow (energy transfer as heat) , kg/s = J/kg
m
Therefore, Equation (2.103) becomes
U + 1 u21 + gz + P V  + Q = U + 1 u22 + gz + P V  + W
 1 2 1 1 1
  2 2 2 2 2
 s

Rearranging gives
U + 1 u22 + gz + P V  – U + 1 u21 + gz + P V  = Q – W
 2 2 2 2 2
  1 2 1 1 1
 s
1 2 2
[U2 – U1] + 2 [u2 – u1] + g (z2 – z1) + (P2V2 – P1V1) = Q – Ws
2 2
∆U = U2 – U1, ∆u = u2 – u1 , (z2 – z1) = ∆z and ∆ (PV) = P2V2 – P1V1
Chemical Engineering Thermodynamics - I 2.31 First Law of Thermodynamics

With these terms, the above equation becomes


1
∆U + 2 ∆u2 + g ∆z + ∆ (PV) = Q – Ws

1
∆U + ∆ (PV) + 2 ∆u2 + g ∆z = Q – Ws ... (2.104)

[Q is heat transferred to the flow system and W is the work done by the system.]

We have, H = U + PV

The change in enthalpy is given by

∆H = ∆U + ∆ (PV) ... (2.105)

Substituting Equation (2.105) in Equation (2.104), we get


1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... (2.106)

This equation is the mathematical statement of the first law of thermodynamics for a steady-
state flow process and is of basic importance in engineering.

Equation (2.106) is often called the first law steady-flow equation.

Each term in Equation (2.106) represents energy per unit mass of the flowing fluid and thus
having the units of J/kg in the SI system.

This equation is applicable to any flow process - reversible as well as irreversible.

Units of g ∆z ⇒ (m/s2) (m) ⇒ s2 ⇒ kg·s2 ⇒  s2  kg ⇒ kg ⇒ kg


m2 kg·m2 kg·m m N·m J
  

Units of 2 ∆u2 ⇒  s  ⇒ kg  s2  ⇒  s2  kg ⇒ kg ⇒ kg


1 m2 kg m2 kg·m m N·m J
      

Units of ∆H, Q and W ⇒ J/kg

∆H = Specific enthalpy change - enthalpy change associated with a unit mass of fluid (J/kg)

∆U = Specific internal energy change in J/kg since U and H in the above equations are the
specific internal energy and specific enthalpy of the fluid, respectively.

In many engineering flow problems, all the energy effects appearing in Equation (2.106) are
not involved. Some processes of interest involving certain restrictions are given in Table 2.1.
Chemical Engineering Thermodynamics - I 2.32 First Law of Thermodynamics

Table 2.1 : Restricted forms of steady-state flow equation


Condition of flow Form of Equation Example of use
1. Adiabatic + external work 1 General flow of fluids
– Ws = ∆H + 2 ∆u2 + g ∆z
effect
2. Adiabatic + external work – Ws = ∆H Adiabatic steam turbine
effect + negligible KE and
PE effects
3. Adiabatic + no external 1 Adiabatic flow nozzle
– ∆H = 2 ∆u2
work effect + negligible
PE effects
4. Adiabatic + no external ∆H = 0 Joule-Thomson expansion
work effect + negligible
KE and PE effects
5. No external work effects Q = ∆H Heat exchanger
+ negligible KE and PE (Heaters and coolers)
effects
Whenever the kinetic energy and potential energy terms are very small compared with the
other terms, they may be neglected. Therefore, in such cases, Equation (2.106) reduces to
∆H = Q – Ws ... (2.107)

This equation of the first law of thermodynamics for a steady-state flow process is analogous
(similar) to Equation (2.9) for a non-flow process.
We have, H = U + PV
In differential form, it becomes
dH = dU + P dV + V dP
dU = dQ – P dV ... P dV - work done in a reversible process
∴ dH = dQ + V dP
Integrating, we get
P2

∆H = Q + ⌠
⌡ V dP ... (2.108)
P1

Substituting Equation (2.108) in Equation (2.106), we get


P2
∆u 2
–Ws = ⌠
⌡ V dP + 2 + g ∆z ... (2.109)
P1
Chemical Engineering Thermodynamics - I 2.33 First Law of Thermodynamics

If the changes in kinetic energy and potential energy are very negligible, then the above
equation reduces to
P2

Ws = – ⌠
⌡ V dP ... (2.110)
P1

This work is called the reversible shaft, flow or pump work of a steady-state flow process
(or simply the work of a steady-flow process).
The work done by the system in non-flow process is
P2


⌡ P dV ... reversible work
P1

The work done by the system in a steady-state flow process is


P2

–⌠
⌡ V dP ... reversible work
P1

In both the cases, V is the volume per unit mass.


Graphical representations of the integral ⌠
⌡ P dV and – ⌠
⌡ V dP are shown in Fig. 2.12.

PdV 1

- VdP
P P
2

V V
(a) Work done in non-flow process (b) Work done in flow process
Fig. 2.12 : Representation of work on P-V diagram
The work done in a non-flow process is given by
P2


⌡ P dV = Q – ∆U (since ∆U = Q – W)
P1

For an isothermal process : ∆U = 0 and ∆H = 0.


Therefore, for an isothermal process (for which ∆U = 0), the above equation becomes
P2


⌡ P dV = Q ... non-flow isothermal process ... (2.111)
P1
Chemical Engineering Thermodynamics - I 2.34 First Law of Thermodynamics

P2

We have, ∆H = Q + ⌠
⌡ V dP
P1

The work done in a steady-state flow process is given by


P2

–⌠
⌡ V dP = Q – ∆H
P1

Thus, for an isothermal process (for which ∆H = 0), it reduces to


P2

–⌠
⌡ V dP = Q ... (2.112)
P1

Therefore, from Equations (2.111) and (2.112), we get


P2 P2


⌡ P dV = – ⌠
⌡ V dP ... for an isothermal process only ... (2.113)
P1 P1

Equation (2.113) indicates that if we represent these two integrals on a P-V diagram, the area
under both the curves is the same / is identical for an isothermal process.
A graphical representation of the work done by the system in isothermal non-flow and
isothermal flow processes is shown in Fig. 2.13.

1 1
5 5

P P
4 2 4 2
3 3

6 V 7 6 V 7

(a) Isothermal non-flow process (b) Isothermal flow process


Fig. 2.13 : Representation of work done on a P-V diagram
Rectangle 1-3-4-5 in (b) = Rectangle 3-2-7-6
For (a), Area under the curve = Area bounded by 1-2-3 + Area of rectangle 3-2-7-6
For (b), Area under the curve = Area bounded by 1-2-3 + Area of rectangle 1-3-4-5
= Area bounded by 1-2-3 + Area of rectangle 3-2-7-6
Chemical Engineering Thermodynamics - I 2.35 First Law of Thermodynamics

∴ The areas are identical.


• Steady flow constant pressure process,
P2

Ws = – ⌠
⌡ V dP = 0 ... [since dP = 0] ... (2.114)
P1

• Steady flow constant volume process,


P2

Ws = – ⌠
⌡ V dP = – V (P2 – P1) ... (2.115)
P1

• Steady flow constant temperature process,


2

Ws = – ⌠
⌡ V dP
1

The constant temperature process is represented by the equation


PV = C = P1V1 = P2V2
P2 2

Ws = – ⌠
C P1 P1
⌡ V dP = – ⌠
⌡ P dP = C ln P2 = P1V1 ln P2 ... (2.116)
P1 1

APPLICATIONS OF THE FIRST LAW STEADY-FLOW EQUATION


Engineering systems that are commonly used may be grouped as :
(i) Work producing systems - heat engines and turbines.
(ii) Work consuming systems - pumps, compressors and refrigerators.
(iii) Neither work producing nor work consuming system - nozzles and heat exchangers.
When a flow process is accompanied by a reduction in pressure, it is called an expansion
process.
An expansion process (flow through nozzles and turbines) results in a decrease in pressure,
whereas a compression process (flow through compressors, pumps and blowers) results in an
increase of pressure. Refer to Table 2.1 for flow equations.
COMPRESSORS
Compression of gases is an important step in many industrial processes, which may be
accomplished using reciprocating compressors or centrifugal compressors. In a compressor,
a definite quantity of gas / air is taken at atmospheric pressure and delivered at any desired
pressure. Reciprocating compressors are capable of developing very high pressures, whereas
centrifugal compressors are capable of handling high volume flow at moderate pressures.
Chemical Engineering Thermodynamics - I 2.36 First Law of Thermodynamics

Compressors are power (work) consuming devices, i.e., they require power / work input.

2 2

Ws

1 1

Fig. 2.14 : Flow through a compressor


The first law of thermodynamics for a steady-state flow process is
∆u2
∆H + 2 + g ∆z = Q – Ws ... Equation (2.106)

If KE and PE changes are presumed to be negligible, then Equation (2.106) reduces to


∆H = Q – Ws ... (2.117)

For adiabatic compression (for an adiabatic compressor - a compressor which is well


insulated) Q = 0 and Equation (2.117) thus becomes
–Ws = ∆H ... (2.118)
–Ws or (–Ws) = Shaft work required for compression

or work input for compression


Equation (2.118) states that the enthalpy of the gas in the compression increases by the
amount of work input.
A reversible process is isentropic if Q = 0, that is, a reversible adiabatic process is isentropic
(constant entropy process). If the gas in the compressor undergoes a compression process which
is reversible and adiabatic then the compression process is isentropic (process occurring at
constant entropy - idealized condition).
For an isentropic compression - reversible adiabatic compression - Equation (2.118) can be
written as
– Ws (isentropic) = ∆Hs ... (2.119)

The isentropic work is the minimum shaft work required for compressing a gas from a given
initial state (conditions) to a given discharge pressure. The shaft work given by Equation (2.119)
is the minimum work, that can be required for an adiabatic compressor with given inlet and outlet
conditions. Actual compressors require more work since the actual process of compression is
irreversible.
Chemical Engineering Thermodynamics - I 2.37 First Law of Thermodynamics

The performance of compressors, pumps and turbines is expressed in terms of isentropic


efficiency of the machine/device. With this the actual performance of the device is compared
with that of an isentropic device for the same inlet and outlet conditions.
The isentropic efficiency of a compressor or simply the efficiency of a compressor / the
compressor efficiency is defined as
Work required by the compressor if it were isentropic
η = Work required by the actual compressor
–Ws (isentropic) ∆Hs
= – Ws = ... (2.120)
∆H
The values of compressor efficiencies usually range from 70 to 80 percent.
Let us obtain expressions for the shaft work required (the work of compression) for an ideal
gas undergoing reversible adiabatic compression and a reversible isothermal compression.
1. Reversible adiabatic compression : For an ideal gas undergoing an isentropic
compression - a reversible adiabatic compression from a pressure P1 to a pressure P2 the shaft
work required can be obtained by using
P2

Ws = – ⌠
⌡ V dP ... (2.121)
P1

For an ideal gas undergoing a reversible adiabatic compression, we have


γ
PV = C ... (2.122)
γ γ γ
∴ P1V1 = PV = P2V2
From this, we can write
P11/γ
V = V1 P ... (2.123)
 
Replacing V from Equation (2.121) using V from Equation (2.123), we get
P2

Ws = – ⌠
P11/γ ... (2.124)
⌡ V1  P  dP
P1
P2 P2

⌠ dP
⌡ P1/γ = – V1P1 ⌠
1/γ 1/γ
Ws = – V1P1 ⌡ P–1/γ dP
P1 P1

P2
1/γ
P  –1/γ + 1
Ws = – V1P1  
– 1/γ + 1P
1

1/γ
– V1P1 γ–1/γ γ–1/γ
= P
(γ – 1) / γ 2 [ – P1 ] … (2.125)
Chemical Engineering Thermodynamics - I 2.38 First Law of Thermodynamics

γ–1/γ
Dividing and multiplying the R.H.S. by P1 gives
1/γ γ–1/γ
– V1 · P1 · P1
Ws =
(γ – 1)/γ · P1
γ–1/γ [Pγ2
–1/γ γ–1/γ
– P1 ] … (2.126)

–γ 1/γ γ–1/γ P2γ–1/γ – P1γ–1/γ


=
(γ – 1)
V1 P1 · P1  γ–1/γ 
 P1 
1 γ–1
γ P2γ–1/γ  γ+ γ (1+γ–1) / γ
= – · V1P1 P  – 1 , ... (P1) = (P1) = P1
(γ – 1)  1 
γ P1V1  P2γ–1/γ
∴ Ws = 1 – P   ... (2.127)
(γγ – 1)   1 
The value of Ws calculated using Equation (2.127) comes out to be –ve which indicates that
work is required for the compression of gas.
The work required for compressing an ideal gas may also be expressed in terms of the initial
and final temperatures (T1 and T2).
For an ideal gas, ∆H = Cp (T2 – T1) ... for any process ... (2.128)
dH = Cp dT

∴ ∆H = ⌠
⌡ Cp dT ⇒ ∆H = Cp ∆T if Cp is constant

We have, –Ws = ∆H
∴ Ws = –∆H ... Equation (2.118)
Combining Equations (2.118) and (2.128), we get
Ws = – Cp (T2 – T1) = Cp (T1 – T2) ... (2.129)
REVERSIBLE ISOTHERMAL COMPRESSION
When cooling during compression is perfect, the temperature remains constant and thus the
compression process is isothermal.
P2

We have, Ws = – ⌠
⌡ V dP
P1

For an ideal gas, PV = RT


RT
∴ V = P
Replacing V by RT/P in the equation for Ws gives
P2
RT
Ws = – ⌠
⌡ P dP ... (2.130)
P1
Chemical Engineering Thermodynamics - I 2.39 First Law of Thermodynamics

Integrating with the limits, we get


P2
Ws = – RT ln P 
 1
P1 V2
Ws = RT ln P  = RT ln V  ... (2.131)
 2  1
since P V = P V = RT, we have P1 = V2
 V1
 1 1 2 2 P2
Note that for a given compression ratio and suction conditions, the work required for
compressing a gas in isothermal compression is less than that required for adiabatic compression.
This is one reason for necessity of cooling in compressors.
Note that if γ = 1, the equations for isothermal and for adiabatic compression are identical.
P2
dP
For a reversible isothermal compression ⇒ Ws = – RT ⌠
⌡ P ... Equation (2.130)
P1

P2
dP
For a reversible adiabatic compression ⇒ Ws = – V1P1 ⌠
1/γ
⌡ 1/γ ... Equation (2.125)
P1 P

γ γ
We have, P1V1 = PV
1/γ 1/γ
P1 V1 = P V
1/γ
1/γ P V
∴ P1 = V1
1/γ
Substituting for P1 , we get
P2 P2
1/γ
– V1 (P V) dP dP
Ws = ⌠
⌡ 1/γ = –P V ⌠
1/γ
⌡ 1/γ
V1
P1 P P1 P

For γ = 1, the above equation reduces to


P2
dP
Ws = – PV ⌠
⌡ P
P1

P2
dP
= – RT ⌠
⌡ P , (since PV = RT) ... same as Equation (2.130)
P1

The work required in a multi-stage compressor is equal to the sum of the work required for
each individual stage. Thus, for a two-stage compressor in which an ideal gas is compressed
reversibly and adiabatically from an initial pressure of P1 to a pressure P' in the first stage, cooled
Chemical Engineering Thermodynamics - I 2.40 First Law of Thermodynamics

to the initial temperature T1 before fed to the second stage and compressed in it from P' to a
pressure P2, the total work required is given by
γRT1  P' γ–1/ γ
1 – P 
P2γ–1/ γ
Ws =  + 1 – P'
   ... (2.132)
(γγ – 1)   1 
For a n-stage compressor with the same compression ratio (r), the total work required is
given by
nγγ γ–1/γ
Ws =
(γγ – 1)
[
RT1 1 – (r) ] ... (2.133)

Here it is assumed that the temperature of the gas is brought down to the initial value
[by interstage cooling arrangements (coolers)] after each stage, i.e., before the gas enters every
next stage.
The reversible isothermal compression and reversible adiabatic compression of an ideal gas
from an initial pressure P1 to a final pressure P2 can be represented on a P-V diagram as shown in
Fig. 2.15.

P2 3 2
4 Adiabatic compression

Isothermal compression

P1 5 1

V
Fig. 2.15 : Reversible isothermal and adiabatic
compression on a P-V diagram (single stage)
The work required in each path Ws = – ⌠
⌡ V dP , i.e., the value of the integral for each

path can be evaluated graphically and is equal to the area enclosed by the curve (either the curve
1-2 or the curve 1-3), the pressure axis and the two horizontal lines at initial and final pressures
(i.e., at P1 and P2). The area 1-2-4-5 represents the work required in compressing the gas
isentropically - reversibly and adiabatically from P1 to P2 in a single stage. The area 1-3-4-5
represents the work required for an isothermal compression. It is clear from Fig. 2.15 that an
Chemical Engineering Thermodynamics - I 2.41 First Law of Thermodynamics

isothermal compression requires less work compared to the work required for an adiabatic
compression. The work required in a reversible isothermal compression (refers to the ideal
condition) is a minimum while that in an adiabatic compression is maximum. However, the
isothermal compression can be approached in practice by compressing the gas in a multistage
compressor with interstage cooling at constant pressure.

P2
Adiabatic compression
(single stage)

Three stage compression


with interstage cooling
P

Isothermal compression
(single stage)

P1

Fig. 2.16 : Three-stage compression with interstage cooling


In an adiabatic compression, the temperature of the gas rises. This increase in temperature
means an increase in the internal energy and since energy must have come from the input energy
to the machine, and consequently this increases the input work requirement. In a multi-stage
compressor, the gas is cooled to its initial temperature between the stages in intercoolers (before
it enters each stage). Because of cooling, the final delivery temperature is reduced. The reduction
of temperature means reduction of the internal energy of the gas delivered. Since this energy
must have come from the energy input to the compressor, there is a decrease in work input for a
given quantity of gas.
Thus, a multi-stage compression, with interstage cooling, approaches more closely the
isothermal compression. The net saving in energy (work input) with a multi-stage compressor
(e.g., a 3-stage compressor) is shown as the shaded area in Fig. 2.16.
TURBINES
Compressors require power input, whereas turbines give positive power output. A turbine is
a work producing device.

The first law equation for a steady-state flow process is


∆u2
∆H + 2 + g ∆z = Q – Ws ... Equation (2.106)
Chemical Engineering Thermodynamics - I 2.42 First Law of Thermodynamics

1 1

Ws

2 2

Fig. 2.17 : Flow through a turbine


For a well insulated turbine with a little change in elevation and with relatively low fluid
 ∆u2 
velocities Q = 0‚ g ∆z = 0 and 2 = 0 , Equation (2.106) reduces to
 
–Ws = ∆H ⇒ Ws = – ∆H = – (H2 – H1) = H1 – H2 ... (2.134)
That is, the work is done by the fluid at the expense of its enthalpy. Ws represents the work
extracted / work produced / work done by the system.
In Ws = – ∆H, – ∆H is a positive quantity and thus Ws represents the shaft work delivered /
produced.
Usually we know P1, T1 and P2. Therefore, we know only H1. To calculate Ws, we must know
H2. For this, we must know the final state of the fluid; which need an additional equation.
A reversible process is isentropic when Q = 0, that is a reversible adiabatic process is
isentropic (constant entropy process ⇒ S2 = S1). If the fluid in the turbine undergoes an
expansion process which is reversible and adiabatic, then the expansion process is isentropic.
With S2 = S1, one can determine the final state and thus H2. For a reversible adiabatic expansion
process, the work done by the fluid is called the isentropic work.
Ws (isentropic) = – ∆Hs ... (2.135)
The shaft work given by Equation (2.135) is the maximum work which can be produced from
an adiabatic turbine for given inlet conditions and a given discharge pressure. [since we assumed
the process of expansion in the turbine to be reversible (ideal case).]
Since the actual expansion is irreversible, actual turbines produce less amount of work
(than that can be produced by a reversible adiabatic turbine).
The isentropic efficiency or simply the efficiency of a turbine is defined by
Work output of the actual turbine
η = Work output of the turbine‚ if it were isentropic
Ws
= W
s (isentropic)
– ∆H H1 – H2
η = = ... (2.136)
– ∆ Hs H1 – H2'
where H1, the enthalpy of the fluid at the inlet of the turbine, H2 at the exit of the actual turbine
and H' at the exit of the turbine, if it were isentropic.
2

The values of η of the turbines range from 70 to 80 percent.


Chemical Engineering Thermodynamics - I 2.43 First Law of Thermodynamics

PUMPS
Pumps are used for transportation of liquids from one vessel to another or through long
pipes.
Pumps required power input - these are work consuming devices - same as compressors.
Therefore, the equation given for adiabatic compressors is applicable to adiabatic pumps.
–Ws = ∆H = (H2 – H1) ... Equation (2.118)
–Ws represents the shaft work required. Therefore, in pumps the enthalpy of a fluid increases
by the amount of work input. In –Ws = ∆H, ∆H is a positive quantity and – Ws thus represents
the shaft work required.
In pumps, work is done on the fluid and this to be provided from an outside source.
For an isentropic process, Equation (2.118) can be written as
– Ws (isentropic) = ∆Hs ... (2.119)
when ∆Hs is the change in enthalpy at constant entropy.
For a reversible adiabatic process (i.e., for an isentropic process), we have
dH = V dP, (since dH = T dS + V dP and dS = 0 for a constant
entropy process)
P2

or ∆H = ⌠
⌡ V dP ... for constant S ... (2.137)
P1

Combining Equations (2.119) and (2.137), we get


P2

–Ws (isentropic) = ∆Hs = ⌠


⌡ V dP ... (2.138)
P1

Usually, for liquids V is independent of P. Therefore, integration of Equation (2.138) gives


–Ws (isentropic) = ∆Hs = V (P2 – P1) ... (2.139)
–Ws represents the shaft work required for pumping.
The isentropic efficiency or simply efficiency of a pump is defined as the ratio of the work
required by the pump, if it were isentropic to the work required by the actual pump for the same
inlet and exit pressures of the fluid.
Work (power) required by the pump‚ if it were isentropic
η = Work (power) required by the actual pump
– Ws (isentropic)
= – Ws
∆ Hs
η = ... (2.140)
∆H
(If Ws = – 400 kJ/kg, it means that the work required is 400 kJ/kg. This can be written as
–Ws = 400 kJ/kg, where –Ws represents the work required. Ws = 500 kJ/kg indicates that the
work equal to 500 kJ/kg is produced / developed, since Ws represents the work produced or
developed.)
Chemical Engineering Thermodynamics - I 2.44 First Law of Thermodynamics

NOZZLES
• A nozzle is primarily used to increase the velocity of a fluid.
• A nozzle is a device which increases the velocity or the kinetic energy of a fluid at the
expense of its pressure drop.
• In a nozzle, the enthalpy of a fluid is converted into kinetic energy and the fluid leaves
the nozzle with increased velocity.
1 2

1 2
Fig. 2.18 : Converging - diverging nozzle
The first law of thermodynamics for a steady-state flow process is given by
∆u2
∆H + 2 + g ∆z = Q – Ws ... Equation (2.106)

Assume that (i) the nozzle is horizontal so that the change in potential energy is zero
(g ∆z = 0), (ii) there is no shaft work, so that Ws = 0, (iii) the nozzle is adiabatic, so that
Q = 0.
With these assumptions, Equation (2.106) reduces to
∆u2
∆H + 2 = 0
2 2
u2 – u1
or (H2 – H1) + 2 = 0 ... (2.141)

In the differential form it is


dH = – u du ... (2.142)
If the inlet velocity (u1) is small compared to the exit velocity (u2), Equation (2.141) becomes
2
u2
H2 – H1 + 2 = 0
2
u2
2 = H1 – H2
∴ u2 = 2 (H1 – H2), m/s ... (2.143)
where (H2 – H1) is in J/kg.
Chemical Engineering Thermodynamics - I 2.45 First Law of Thermodynamics

Let us derive an expression for the exit velocity if an ideal gas at P1 and T1 enters a reversible
adiabatic nozzle with negligible velocity.
We have, dH = – u du ... Equation (2.142)
For a reversible process, dH = dQ + V dP
(since dH = dU + P dV + V dP and dQ = dU + P dV)
For a reversible adiabatic process, it becomes
dH = V dP ... (2.144)
Combining Equations (2.142) and (2.144), we get
– u du = V dP ... (2.145)
For an ideal gas undergoing a reversible adiabatic process, we have
γ
PV = C
γ
∴ P1Vγ1 = PV = C

V =
P11/γ V
P 1

Substituting this value of V in Equation (2.145), it becomes


1/γ dP
– u du = (P1) V1 × 1/γ
P

u du can be written as d  2  .
u2
 

– d  2  = V1(P1)
u2 1/γ dP
  P
1/γ

Integrating between the limits : inlet velocity = u1 = 0 (assumed negligible) and


inlet pressure = P1; and exit velocity = u2 and exit pressure = P2, we get
u2 P2 P2

u 
2 dP
–⌠
⌡ d  2  = V1 (P1) ⌠ ⌡ 1/γ = V1 (P1) ⌠
1/γ 1/γ –1/γ
⌡P dP
0 P P 1 P 1

P2
2
u2 1/γ
 P–1/γ +1 
– 2 = V1 (P1)  
– 1/γ + 1P
1

2 1/γ
u2 – V1 (P1)
2 =
(γ – 1) / γ [Pγ 2
–1/γ γ–1/γ
– P1 ]
1/γ γ–1/γ
– V1 P1 · P1 P2γ–1/γ 
= P  – 1
(γ – 1) / γ  1 
Chemical Engineering Thermodynamics - I 2.46 First Law of Thermodynamics

2
u2 γP1V1  P2γ–1/γ
= 1 –   
(γ – 1)  P1
2 
 1 γ–1 
 γ+ γ (1+ γ–1) /γ γ /γ 
1/γ γ–1/γ
since P1 · P1 = (P1) = (P1) = (P1) = P1
1/2
 2γγP1V1  P2γ–1/γγ 
∴ u2 =  1 – P   ... (2.146)
 γ
(γ – 1)   1  
For an ideal gas, P1V1 = RT1
Therefore, Equation (2.146) becomes
1/2
 2γγRT1  P2γ–1/γγ 
u2 =  1 – P   ... (2.147)
 γ
(γ – 1)   1  
For an ideal gas, we have
Cp – Cv = R ... (2.148)
Cp
We have, Cv = γ
Cp
Therefore, Cv =
γ
Substituting for Cv, Equation (2.148) becomes
Cp
Cp – = R
γ
γ – 1 = R
Cp  
 γ 
γR
∴ Cp = ... (2.149)
(γ – 1)
Substituting Equation (2.149) in Equation (2.147), we get
1/2
  P2γ–1/γγ 
u2 =  2Cp T1 1 – P   ... (2.150)
   1  
THROTTLING PROCESS (JOULE - THOMSON EXPANSION)
When a fluid flows through a constriction, such as a partially opened valve, an orifice, or a
porous plug, there is an appreciable drop / reduction in the pressure of the fluid, and the flow is
said to be throttled.
Throttling process is a steady-state flow process in which a fluid flowing through a
constriction, such as a partially opened valve, an orifice, or a porous plug, undergoes an
appreciable drop in pressure. The throttling process may be considered as an adiabatic, since the
expansion takes place in a very short time over a very small region, so that the heat transfer
between the system and its surrounding is negligible.
Chemical Engineering Thermodynamics - I 2.47 First Law of Thermodynamics

Throttling process normally encountered in actual practice was investigated by Joule and
Thomson. In their famous porous-plug experiment, a gas was allowed to flow steadily through a
porous plug to measure the changes in temperature and pressure of the gas. The porous-plug
apparatus consists of a long horizontal and insulated tube containing a porous plug of silk or
cotton (a throttle) at the centre of the tube. In order to measure the temperatures and pressures of
the gas on either side of the porous plug, i.e., at upstream and downstream of the plug, provisions
are made as shown in Fig. 2.19.
Porous plug

T1 T2 Thermometer

Gas

P1 P2

Fig. 2.19 : Porous-plug apparatus (Joule - Thomson expansion)


The first law equation for a steady-state process is
∆u2
∆H + g ∆z + 2 = Q – Ws ... Equation (2.106)

The tube is insulated to ensure adiabatic conditions, so Q = 0.


There is no shaft work involved in throttling, so Ws = 0.

∆u2
Since the tube is horizontal, g ∆z = 0. There is no significant change in K.E., so 2 = 0.

With these values, Equation (2.106) reduces to


∆H = 0
or H2 – H1 = 0 ⇒ H2 = H1 ... (2.151)

That is, whenever a fluid flows steadily from a high pressure to a low pressure through a
constriction, such as a porous plug or a partially opened valve, the enthalpy of the fluid remains
constant.
The enthalpy of a fluid in an adiabatic throttling process remains constant - the enthalpy of
the fluid after throttling is equal to the enthalpy of the fluid before throttling. Therefore,
a throttling process is an isenthalpic process. (It is a constant enthalpy process.)
Since the enthalpy of an ideal gas is a function of temperature only (i.e., it depends on
temperature only), a throttling process does not change the temperature of an ideal gas - no
change in temperature occurs on throttling an ideal gas.
Chemical Engineering Thermodynamics - I 2.48 First Law of Thermodynamics

For an ideal gas,


T2

∆H = ⌠
⌡ Cp dT = Cp (T2 – T1)
T1

... valid for any process involving ideal gas


For a throttling process, ∆H = 0
Therefore, for an ideal gas undergoing a throttling process,
T2

∆H = ⌠
⌡ Cp dT = Cp (T2 – T1) = 0
T1

∴ T2 = T1
That is, if the fluid in a throttling process is an ideal gas, no change in temperature occurs.
The internal energy of an ideal gas is a function of temperature only. Thus, for an ideal gas,
the throttling process takes place at constant enthalpy, constant internal energy and at constant
temperature.
For an ideal gas, U = f (T) only, i.e., U is a function of T only
The enthalpy of a fluid is given by
H = U + PV
For an ideal gas, PV = RT
Substituting for PV by RT, the above equation becomes
H = U + RT
Since U is a function of T only and H = U + RT, H is a function of temperature only.
For a finite process, the change in H is given by
∆H = ∆U + ∆ (PV)
For a throttling process with any fluid, we have
∆H = 0
Therefore, ∆H = ∆U + ∆ (PV) = 0
∆U + ∆ (PV) = 0 ... (2.152)
If the fluid is an ideal gas, then during the throttling process, temperature (T) does not
change (i.e., T is constant).
For an ideal gas, PV = RT
∴ ∆ (PV) = 0 since PV is constant as R and T are constants
Combining this with Equation (2.152), we get
∆U + ∆ (PV) = 0
∆U + 0 = 0
∆U = 0 ⇒ U2 = U1 ... (2.153)
That is, the internal energy of an ideal gas remains constant in a throttling process.
Chemical Engineering Thermodynamics - I 2.49 First Law of Thermodynamics

For an ideal gas undergoing a throttling process, T2 = T1. However, for most real gases at
moderate conditions of T and P, a reduction in pressure at constant enthalpy results in a decrease
in temperature (i.e., if a real gas is allowed to expand through a porous plug under adiabatic
conditions, it gets cooled appreciably).
The phenomenon of change of temperature resulted when a gas is allowed to expand
adiabatically from a high pressure to a low pressure is known as the Joule - Thomson effect. The
Joule - Thomson effect is zero for an ideal gas (since no change in T occurs).
In the throttling process, the enthalpy of the system remains constant and if the system is an
ideal gas, the internal energy of the system also remains constant.
For an ideal gas, the throttling process takes place with constant H, constant U and constant
T. In this expansion process, there are changes in P and V. So the internal energy of an ideal gas,
at constant temperature is independent of the pressure and volume. This can be represented
mathematically as follows :
Dividing Equation (2.153) by ∆P, the pressure drop involved in the porous-plug experiment,
we get
∆U = 0
 ∆P  ... (2.154)
 T
In this equation, the subscript T is added to the parenthesis as the temperature does not
change.
Equation (2.154) is applicable for a finite pressure drop. For a differential pressure drop,
it may be rewritten as
∂U = 0
 ∂P  ... (2.155)
 T
∆U
Similarly, ∆V = 0 ... for a finite volume change
 T
∂U = 0
∂V ... for a differential change in volume ... (2.156)
 T
The internal energy of an ideal gas is independent of the pressure and volume, and is a
function of temperature only.
We have, H = U + PV
Differentiating this equation with respect to volume at constant temperature, we get
∂H = ∂U + ∂(PV)
∂V     ... (2.157)
 T ∂VT  ∂V T
∂U
For an ideal gas,   is zero. With this, Equation (2.157) becomes
∂VT
∂H = ∂(PV)
∂V   ... (2.158)
 T  ∂V T
Chemical Engineering Thermodynamics - I 2.50 First Law of Thermodynamics

For an ideal gas, PV is equal to RT and hence is constant.


∂(PV) is zero and hence
As PV is constant, its differential, i.e.,  
 ∂V T
∂H = 0
∂V ... for an ideal gas ... (2.159)
 T
Similarly, it can be proved that
∂H
 ∂P  = 0 ... for an ideal gas ... (2.160)
 T
Thus, it follows from the above equations that the enthalpy of an ideal gas is independent of
the volume and pressure and is a function of temperature only.
Note that for real gases :
∂U ≠ 0 , ∂U ≠ 0
∂V  ∂P 
 T  T
∂H ∂H
∂V ≠ 0 ,  ∂P  ≠ 0
 T  T
That is, the enthalpy and internal energy of a real gas are not functions of temperature alone.
The throttling process - Joule-Thomson expansion of a real gas takes place not with constant
internal energy but with constant enthalpy.
µJT)
JOULE - THOMSON COEFFICIENT (µ
Since H is a state function (determined by the thermodynamic state of the system), dH is a
complete differential. Thus, if H is a function of P and T, i.e., H = f (P, T), then the total
differential of H is
∂H ∂H
dH =   dP +   dT ... (2.161)
 ∂P T  ∂T P
For a throttling process, dH = 0. Therefore, Equation (2.161) becomes
∂H ∂H
 ∂P  dP +  ∂T  dT = 0
 T  P
Rearranging gives
∂H dT = – ∂H dP
 ∂T   ∂P  ... at constant H
 P  T
dT (∂H/∂P)T
Therefore, dP = – ... at constant H
(∂H/∂T)P
∂T (∂H/∂P)T
or ∂P = – (∂H/∂T) ... (2.162)
 H P

(∂H/∂T)P is the heat capacity of the gas at constant pressure, i.e., (Cp).
Chemical Engineering Thermodynamics - I 2.51 First Law of Thermodynamics

Equation (2.162) thus becomes


∂T = – 1 ∂H
∂P Cp  ∂P T ... (2.163)
 H
∂T
The quantity   is called the Joule-Thomson coefficient and it is denoted by the symbol
∂PH
µJT. It is equal to the rate of change of temperature with the pressure in a throttling process
(a fluid-flow process through a porous plug or throttle).
1 ∂H
µJT = – C   ... (2.164)
p  ∂P T

Joule-Thomson Coefficient of an Ideal Gas :


1 ∂H
We have, µJT = – C   ... Equation (2.164)
p  ∂P T
We have, H = U + PV
Differentiating the above equation with respect to P keeping T constant gives
∂H = ∂U + ∂(PV)
 ∂P      ... (2.165)
 T  ∂P T  ∂P T
∂H
Substituting for   , Equation (2.164) becomes
 ∂P T
1  ∂U ∂(PV) 
µJT = – C    +  ∂P  ... (2.166)
p   ∂P T  T 
∂U
Since for an ideal gas   is zero, as seen earlier, and since PV = RT, PV is constant at
 ∂P T
∂(PV)
constant temperature, it follows that   is also zero.
 ∂P T
Equation (2.166) reduces to
µJT = 0 ... for an ideal gas
The Joule-Thomson coefficient of an ideal gas is zero. As µJT = 0, there should be no change
of temperature when an ideal gas expands through a porous plug, i.e., a throttle. Note that
Equation (2.166) is quite general and is applicable to any gas.
With real gases in a throttling process, a change in temperature takes place and µJT is positive
at ordinary temperature and low to moderate pressure and there is a fall in temperature - the
temperature of a real gas falls as a result of its throttled expansion. At very high pressure, µJT is
negative and the temperature of a real gas increases as a result of its throttled expansion.
Suppose a series of experiments are conducted on the same gas keeping the same inlet
pressure (P1) and temperature (T1) by varying the downstream/exit pressure (downstream of the
porous plug) to various values P2, P3, P4, etc., we get different exit temperatures as T2, T3, T4, etc.
If we plot a graph of exit T v/s exit P, we get a smooth curve passing through all these points.
Chemical Engineering Thermodynamics - I 2.52 First Law of Thermodynamics

This curve is of constant enthalpy as H1 = H2 = H3 = H4, etc. and is called an isenthalpic curve
or an isenthalpe as shown in Fig. 2.20. By repeating the experiment with new values for initial
pressure and temperature, we can get a series of / a family of isenthalpic curves for the gas as
shown in Fig. 2.20. The curve passing through the maxima of these isenthalpic curves is called
the inversion curve.

Maximum inversion temperature

Isenthalpic curve

mJT = slope = 0

Heating region
Cooling region
mJT +ve, i.e., > 0 mJT -ve, i.e., < 0
T (exit)

Inversion curve
Isenthalpic curve

P (exit)

Fig. 2.20 : Isenthalpic curves and the inversion curve of a real gas
The slope of an isenthalpic curve on a T - P plot at any point is called the Joule - Thomson
coefficient (µJT). The point at which µJT is zero is called the inversion point - and the locus of all
the inversion points is called the inversion curve. The region inside the inversion curve (i.e., to
the left of the inversion curve), where µJT is positive is called the cooling region. In this region,
throttling results in cooling of a real gas (in this region, temperature falls with fall of pressure).
The region outside the inversion curve (i.e., to the right of the inversion curve), where µJT is
negative, is called the heating region. In this region, throttling results in heating of the real gas
(in this region, temperature rises with fall of temperature).
Cooling can take place only if the temperature of the gas before throttling is below the point
at which the inversion curve intersects the temperature axis, i.e., below the maximum inversion
temperature. For nearly all substances (except hydrogen and helium), the maximum inversion
temperature is above the normal ambient temperature and therefore, cooling can be achieved by
the Joule-Thomson effect. Hydrogen and helium gases should be precooled in heat exchangers
below the maximum temperatures before they are throttled. For liquefaction of a gas, it has to be
cooled below its critical temperature. The throttling process finds application in the liquefaction
of gases and in refrigeration systems.
Chemical Engineering Thermodynamics - I 2.53 First Law of Thermodynamics

SOLVED EXAMPLES
Example 2.1 :
An elevator of mass 2500 kg rests at a level of 7.5 m above the base of an elevator shaft. The
elevator is raised to a height of 75 m from the base of the shaft, where the cable holding it
breaks. It falls freely to the base where it is brought to rest by a strong spring.
Calculate : (i) The potential energy of the elevator in its initial position.
(ii) The potential energy of the elevator in its highest position.
(iii) The work done in raising the elevator.
(iv) The kinetic energy and velocity of the elevator just before it strikes the spring.
Given : The acceleration due to gravity is 9.81 m/s2.
Solution :
Mass of the elevator = m = 2500 kg
(i) To find the potential energy of the elevator in its initial position :
z = 7.5 m, g = 9.81 m/s2
P.E. = mgz
= 2500 × 9.81 × 7.5 = 183937.5 J
= 183.9375 kJ ≈ 183.94 kJ ... Ans. (i)
(ii) To find the potential energy of the elevator in its highest position :
z for the highest position = 75 m
P.E. = mgz
= 2500 × 9.81 × 75
= 1839375 J ≈ 1839.4 kJ ... Ans. (ii)
(iii) To find the work done in raising the elevator from a height of 7.5 m to 75 m :
Work done = mg (z2 – z1)
= 2500 × 9.81 × (75 – 7.5)
= 1655437.5 J = 1655.44 kJ ... Ans. (iii)
(iv) To find the velocity of the elevator just before it strikes the spring :
P.E. at the highest position = 1839.4 kJ = 1839.4 × 103 J
P.E. is converted into K.E.
K.E. = P.E. = 1839.4 × 103 J
1
K.E. = 2 mu2 = 1839.4 × 103

∴ u =
2 × 1839.4 × 1031/2 = 38.36 m/s ... Ans. (iv)
 2500 
Chemical Engineering Thermodynamics - I 2.54 First Law of Thermodynamics

Example 2.2 :
1 kg of water in a waterfall is flowing down from a height of 100 m and at the bottom of the
fall it joins a river. Neglecting energy exchange between the water and the surroundings and
assuming the river downstream velocity to be negligible, calculate : (i) the potential energy of
water at the top of the fall, (ii) the kinetic energy just before the water strikes the bottom and
(iii) the change in temperature of water when it enters the river.
Take g = 9.81 m/s2 and specific heat of water = 4.187 kJ/(kg·K)
Solution :
m = 1 kg, z = 100 m and g = 9.81 m/s2
(i) To find the potential energy of water at the top of the fall :
P.E. = mgz
= 1 × 9.81 × 100 = 981 J ... Ans. (i)
(ii) To find the kinetic energy just before the water strikes the bottom :
K.E. = P.E. = 981 J ... Ans. (ii)
(iii) To find the change in temperature of the water when it enters the river :
Here the K.E. of 981 J will appear as heat in the water.
∴ Q = mC ∆T
Q
∴ ∆T = mC , where C = 4.187 kJ/(kg·K) = 4187 J/(kg·K)

981
= = 0.234 K or 0.234°C ... Ans. (iii)
1 × 4187

Example 2.3 :
A ball with a mass of 10 g at 27°C (300 K) is dropped from a height of 10 m. Calculate :
(i) The speed of the ball as it reaches the ground and
(ii) The temperature rise of the ball if all its kinetic energy is transformed into internal
energy as the ball is suddenly stopped after 10 m.
Data : Specific heat of ball material = 125.6 J/(kg·K) and volume change of ball material is
negligible.
Solution :
m = 10 g = 0.01 kg, g = 9.81 m/s2, z = 10 m
P.E. = mgz
= 0.01 × 9.81 × 10 = 0.981 J
Chemical Engineering Thermodynamics - I 2.55 First Law of Thermodynamics

(i) To find the speed of the ball as it reaches ground :


K.E. of the ball as = P.E. of the ball = 0.981 J
 it reaches ground 
1
K.E. = 0.981 = 2 mu2

∴ u =
2 × 0.9811/2 = 14 m/s
 0.01 
Velocity of the ball or speed of the ball = 14 m/s
(ii) To find the temperature of the ball when it is suddenly stopped :
K.E. is converted into Q.
∴ Q = K.E. = P.E. = 0.981 J
∴ Q = mC ∆T = 0.981 J
Q
∴ ∆T = mC

0.981
∴ ∆T = = 0.781 K = 0.78 K or 0.78°C ... Ans.
0.01 × 125.6
Example 2.4 :
Liquid water at 100°C (373 K) and 101.3 kPa has an internal energy of 420 kJ/kg.
The specific volume of liquid water at these conditions is 1.04 × 10–3 m3/kg. The water is
converted to the vapour state at 200°C (473 K) and 700 kPa. At these conditions, its specific
volume is 0.30 m3/kg and enthalpy is 2844 kJ/kg. Calculate :
(ii) The enthalpy of the liquid water and
(ii) The changes in internal energy and enthalpy associated with the vaporization process.
Solution :
(i) Water - Liquid state : T = 373 K, P = 101.3 kPa = 101.3 kN/m2, UL = 420 kJ/kg,
VL = 1.04 × 10–3 m3/kg
The enthalpy of the liquid water per kg in terms of UL , P and VL is given by
HL = UL + PVL
= 420 + 101.3 × 1.04 × 10–3 = 420.10 kJ/kg ... Ans. (i)
(ii) Water - Vapour state : T = 473 K, P = 700 kPa, HV = 2844 kJ/kg, VV = 0.30 m3/kg
The change in enthalpy associated with the vaporization process is
∆HV = HV – hL
= 2844 – 420.10
= 2423.9 kJ/kg ... Ans. (ii)
Chemical Engineering Thermodynamics - I 2.56 First Law of Thermodynamics

The internal energy of water in the vapour state is


HV = UV + PVV
UV = HV – PVV
= 2844 – 700 × 0.30
= 2634 kJ/kg
The change in internal energy associated with the vaporization process is
∆Uvap. = UV – UL ... Process : Water (L) → Water (V)
= 2634 – 420
= 2214 kJ/kg … Ans. (iii)
Example 2.5 :
A sample of 10 g of saturated benzene at 1 atm is vaporized using a 12 V, 0.5 A
electric supply. The boiling point and latent heat of vaporization of benzene are 353.2 K and
30.8 × 103 kJ/kmol respectively. Calculate :
(i) The change in internal energy and
(ii) The time required for complete vaporization.
Solution :
Latent heat of vaporization of benzene = ∆HV = 30.8 × 103 kJ/kmol
Amount of benzene to be converted into the vapour state = 10 g = 0.01 kg
Molecular weight of benzene (C6H6) = 78
0.01
Moles of benzene = n = 78 = 1.28 × 10–4 kmol

∴ Heat required for vaporization = n·∆Hv = 1.28 × 10–4 × 30.8 × 103


∆H = 3.9424 kJ = 3942 J
Electric supply : 12 V and 0.5 A
The electric energy supplied appears as heat to vaporize benzene.
VIt = ∆H
∴ 0.50 × 12 × t = 3942
∴ t = 657 s = 10.95 min … Ans. (ii)
Assume that the molar volume of saturated benzene to be negligible compared to that of
saturated benzene vapour.
∴ ∆V = VV – VL ≈ VV
As vapour behaves ideally, we have
PVV = nRT
Chemical Engineering Thermodynamics - I 2.57 First Law of Thermodynamics

For a constant pressure process,


∆H = ∆U + P ∆V
∴ ∆U = ∆H – P ∆V
= ∆H – PVV
= ∆H – nRT, R = 8.314 kJ/(kmol.K)
= 3.9424 – 1.28 × 10–4 × 8.314 × 10.95
= 3.9307 ≈ 3.931 kJ … Ans. (i)
Example 2.6 :
A cylinder fitted with a piston of volume 0.1 m3 contains 0.5 kg of steam at 500 kPa.
Calculate :
(i) The heat to be supplied to raise the temperature of the steam to 550°C (823 K), keeping
the pressure constant and
(ii) The work done in the process.
Data :
At 500 kPa : VL = 1.093 × 10–3 m3/kg, VV = 0.3749 m3/kg, UL = 639.68 kJ/kg,
UV = 2561.2 kJ/kg, HL = 640.23 kJ/kg, HV = 2748.70 kJ/kg, T = 425 K
At 823 K : HL = 3592.8 kJ/kg, VV = 0.7575 m3/kg
Solution :
Volume = 0.1 m3 and mass of steam = 0.5 kg
0.10
Initial specific volume of steam = V1 = 0.50 = 0.2 m3/kg

Let x represent the quality of steam (dryness fraction of steam).


mV
x = m +m
L V

Evaluate x at 500 kPa :


V1 = xVV + (1 – x) VL
0.20 = x (0.3749) + (1 – x) × 1.093 × 10–3
∴ x = 0.532
H1 = H of steam initially, i.e., at 500 kPa and 425 K
= xHV + (1 – x) HL
= 0.532 (2748.7) + (1 – 0.532) (640.23)
= 1761.94 kJ/kg
H2 = H of saturated steam at 823 K
= 3592.8 kJ/kg
Chemical Engineering Thermodynamics - I 2.58 First Law of Thermodynamics

It is a constant pressure process. Therefore,


Q = m (H2 – H1)
= 0.5 (3592.8 – 1761.94)
= 915.43 kJ
Heat to be supplied = 915.43 kJ ... Ans. (i)
V2 = Final specific volume
= VV at 823 K = 0.7575 m3/kg
Work done in the process per kg of steam,
W = P (V2 – V1)
= 500 (0.7575 – 0.2) = 278.75 kJ/kg
Given : 0.5 kg of steam
∴ Work done, W = 0.50 (278.75) = 139.375 ≈ 139.4 kJ ... Ans. (ii)
[W = mP (V2 – V1) ... as V2 and V1 are in m3/kg]
Example 2.7 :
A steam boiler having a volume of 2.3 m3 initially contains 1.7 m3 liquid water in equilibrium
with 0.6 m3 vapour at 100 kPa. The boiler is heated keeping the inlet and outlet valves closed.
The relief valve is set such that it lifts when the pressure in the boiler reaches 5500 kPa. Find the
amount of steam supplied to the boiler contents before the relief valve lifts.
Data :
At 100 kPa : VL = 1.043 × 10–3 m3/kg, VV = 1.694 m3/kg, UL = 417.36 kJ/kg,
UV = 2506.1 kJ/kg
At 5500 kPa : UV = 2593.4 kJ/kg, VL = 1.3025 × 10–3 m3/kg, VV = 35.94 × 10–3 m3/kg
UL = 1176.62 kJ/kg
Solution :
Hint : It is a constant volume process. For such a process, Q = m (U2 – U1). Therefore,
to find Q, we have to find U1 and U2 and for U1 and U2, we have to find x1 and x2 of steam.
First find the quality of the steam (x1) - dryness fraction at 100 kPa.
Water volume = 1.7 m3, Vapour volume = 0.60 m3
Volume of vapour x1 V V
Volume of water = (1 – x1) VL
0.60 x1 (1.694)
1.7 =
(1 – x1) × 1.043 × 10–3
0.353 (1 – x1) × 1.043 × 10–3 = 1.694 x1
∴ x1 = 2.173 × 10–4
Chemical Engineering Thermodynamics - I 2.59 First Law of Thermodynamics

It is a constant volume process.


Initial internal energy = x U + (1 – x ) U
 of the contents  1 V 1 L

U1 = 2.173 × 10–4 (2506.1) + (1 – 2.173 × 10–4) (417.36)


U1 = 417.81 kJ/kg
Initial specific volume = x V + (1 – x ) V
 of the contents  1 V 1 L

= 2.173 × 10–4 (1.694) + (1 – 2.173 × 10–4) (1.043 × 10–3)


= 1.411 × 10–3 m3/kg
V
Mass of the contents = m = Specific volume , V = 2.3 m3

2.3
= = 1630.05 kg
1.411 × 10–3
At 5500 kPa : To find the final quality of the steam (x2) :
Initial specific volume = Final specific volume as both V and m are constant
∴ 1.411 × 10–3 = x2 (35.94 × 10–3) + (1 – x2) (1.3025 × 10–3)
∴ x2 = 3.13 × 10–3
U2 = Final internal energy of the contents
U2 = x2 UV + (1 – x2) UL
= 3.13 × 10–3 (2593.4) + (1 – 3.13 × 10–3) (1176.62)
= 1181.05 kJ/kg
The amount of heat to be supplied to the contents of boiler is
Q = m (U2 – U1) ... for constant volume process
= 1630.05 (1181.05 - 417.81)
= 1244119.4 kJ
= 1.244 × 106 kJ ... Ans.
Example 2.8 :
It is desired to fill 10 kg wet steam of quality 0.85 at 200°C (473 K) in a tank. Estimate the
size of the tank and the internal energy of steam in the tank.
Given : At 200°C (from steam table) :
Property Saturated liquid Saturated vapour
Internal energy 850.57 kJ/kg 2593.1 kJ/kg
Specific volume 1.1565 × 10–3 m3/kg 0.1272 m3/kg
Chemical Engineering Thermodynamics - I 2.60 First Law of Thermodynamics

Solution :
Hint : We are supplied with the specific volumes of saturated liquid and vapour as well as
mass of steam. Make use of this data through the relation of specific volumes of mixture,
saturated liquid and saturated vapour to get the size, V, of the tank.
Amount of steam = m = 10 kg
Quality of steam = x = 0.85
VL = 1.1565 × 10–3 m3/kg, VV = 0.1272 m3/kg
Let V be the volume of the tank.
(i) To find the size of the tank :

Specific volume of the wet steam = of saturated liquid +  saturated vapour 
Specific volume Specific volume of

V
m = xVV + (1 – x)VL
V
10 = 0.85 (0.1272) + (1 – 0.85) × 1.1565 × 10
–3

∴ V = 1.0829 ≈ 1.083 m3 ... Ans.


Similarly,
Internal energy of the wet steam = xUV + (1 – x) UL
= 0.85 × 2593.1 + (1 – 0.85) × 850.57
= 2331.7 kJ/kg ... Specific internal energy
Given : 10 kg wet steam
∴ Internal energy of 10 kg wet steam = 10 × 2331.7 = 23317 kJ
= 23.317 MJ ≈ 23.32 MJ ... Ans.
Example 2.9 :
A tank having a volume of 1 m3 contains 2.5 kg wet steam at 120°C (393 K). Find the quality
of steam in the tank.
Data : At 120°C :
Specific volume of saturated liquid = 1.061 × 10–3 m3/kg
Specific volume of saturated vapour = 0.8857 m3/kg
Solution :
The relationship between specific volumes of wet steam, saturated liquid and saturated
vapour is
V
m = xVV + (1 – x) VL
where V = Volume of the tank = 1 m3
m = Mass of the steam = 2.5 kg
Chemical Engineering Thermodynamics - I 2.61 First Law of Thermodynamics

VL = 1.061 × 10–3 m3/kg and VV = 0.8857 m3/kg


x = Quality of the steam/dryness fraction of the steam
1
∴ 2.5 = x (0.8857) + (1 – x) × 1.061 × 10
–3

∴ 0.4 = x (0.8857) + (1 – x) × 1.061 × 10–3


0.884839 x = 0.398939
∴ x = 0.4509 ≈ 0.451
Quality of the wet steam = 0.451 ... Ans.
Example 2.10 :
A rigid tank of volume 1 m3 initially contains a mixture of liquid water and vapour at 100°C
(373 K). Energy is transferred as heat to the tank so that the tank contains saturated vapour at
250°C (523 K). Find the mass of steam in the tank. Also find the mass of liquid water and water
vapour initially contained in the tank.
Data : At 100°C : Wet steam of unknown quality.
Specific volume of saturated liquid = 1.0437 × 10–3 m3/kg
Specific volume of saturated vapour = 1.673 m3/kg
At 250°C : Dry saturated steam with x = 1
Specific volume of saturated vapour = 0.05004 m3/kg
Solution :
To find m, make use of specific volume of steam provided.
Volume of the tank = V = 1 m3
During this heating process, the volume of the tank (rigid tank) and the mass of the steam
remains constant.
Hence, the specific volume of the steam remains constant.
∴ Specific volume at 100°C = Specific volume at 250°C
= 0.05004 m3/kg ... Given
Specific volume of steam at 100°C = 0.05004 m3/kg
V
Specific volume = m

where, m = Mass of steam (initially)


1
∴ 0.05004 = m

1
∴ m = 0.05004 = 19.984 kg

Mass of steam at 100oC = 19.984 ... Ans.


Chemical Engineering Thermodynamics - I 2.62 First Law of Thermodynamics

We know the mass of wet steam. To find the masses of liquid water and water vapours in it,
we should know the condition (dryness fraction) of steam. So let us find x (the quality of steam)
at 100°C (which is to be used to find the mass of liquid water and water vapour).
V
At 100°C : m = Specific volume at 100°C = Specific volume at 250°C
= 0.05004
0.05004 = xVV + (1 – x) VL
0.05004 = x (1.673) + (1 – x) (1.0437 × 10–3)
∴ x = 0.0293
∴ Mass of liquid water = (1 – x) (Mass of steam)
= (1 – 0.0293) (19.984)
= 19.398 kg ... Ans.
Mass of water vapours = x · (Mass of steam)
= 0.0293 × 19.984
= 0.5855 kg ... Ans.
Example 2.11 :
At one particular instant a steam boiler contains saturated water vapour at 200°C (473 K).
All the valves of the boiler are closed. After some time, due to energy loss as heat to the
surroundings, the temperature of the boiler drops to 110oC (383 K). Find the ratio of the mass of
liquid to the mass of vapour in the boiler at 110°C (383 K).
Data :
At 200°C : P = 15.549 bar, Specific volume of saturated vapour = 0.1272 m3/kg
Enthalpy of saturated vapour = 2790.9 kJ/kg
At 110°C : P = 1.4327 bar, Specific volume of saturated water = 1.0519 × 10–3 m3/kg
Specific volume of saturated water = 1.21 m3/kg
Enthalpy of saturated water = 481.32 kJ/kg
Enthalpy of saturated vapour = 2691.3 kJ/kg
Solution :
At 200°C : P = 15.549 bar = 15.549 × 105 N/m2 = 15.549 × 102 kN/m2 = 15.549 × 102 kPa
We have : UV = HV – PVV
= 2790.9 – 15.549 × 102 × 0.1272
= 2593.12 kJ/kg
At 110°C : P = 1.4327 bar = 1.4327 × 102 kN/m2 = 1.4327 × 102 kPa
UL = HL – PVL
= 461.32 – 1.4327 × 102 × 1.0519 × 10–3
= 461.17 kJ/kg
UV = HV – PVV
= 2691.3 – 1.4327 × 102 × 1.21 = 2517.94 m3/kg
Chemical Engineering Thermodynamics - I 2.63 First Law of Thermodynamics

Boiler is a closed vessel. Hence, specific volume remains constant during the process of heat
loss (as V of the boiler and mass of the steam are constant/fixed).
Specific volume at 110oC = Specific volume at 200oC = 0.1272 m3/kg ... Given
At 110°C : Specific volume of steam = 01272 m3/kg
Let x be the quality of steam at 110oC.
∴ VS = xVV + (1 – x) VL
0.1272 = x (1.21) + (1 – x) (1.0519 × 10–3)
∴ x = 0.1043
Internal energy of steam at 110°C = US
US = xUV + (1 – x) UL
= 0.1043 (2517.94) + (1 – 0.1043) (461.17)
= 675.67 kJ/kg
The ratio of the mass of liquid water to the mass of water vapour is given by
mL UV – US
mV = US – UL
2517.94 – 675.69
= 675.69 – 461.17
= 8.588 ... Ans.
Example 2.12 :
The following results were obtained by conducting a trial run on steam turbine power plant :
At inlet to boiler At exit of turbine
Mass flow rate = 3600 kg/h Steam velocity = 25 m/s
Enthalpy of water = 850 kJ/kg Elevation from datum level = 0 m
Elevation from datum level = 4.3 m Enthalpy of steam = 2625 kJ/kg
Water velocity = 5 m/s
Find the power developed by the turbine if the heat added in the boiler is 2100 kJ/s.
Solution :
To find the power developed by the turbine, use the first law equation applicable for a
steady-state process to get W in J/kg then convert it in J/s with the help of mass flow rate in kg/s
and then use conversion factor to get the answer in hp.
Mass flow rate of water = 3600 kg/h
3600
= 3600 = 1 kg/s

Heat added in the boiler = 2100 kJ/s


2100 kJ/s
= 1 kg/s = 2100 kJ/kg

Q = 2100 × 103 J/kg


Chemical Engineering Thermodynamics - I 2.64 First Law of Thermodynamics

The first law of thermodynamics for a steady-state flow process is given by


1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... per unit mass of fluid

Each term of this equation has the units of J/kg.


∆H = H2 – H1 = 2625 – 850 = 1775 kJ/kg = 1775 × 103 J/kg
1 1 2 2 1
2 ∆u = 2 (u2 – u1) = 2 [(25) – (5) ] = 300 J/kg
2 2 2

g ∆z = g (z2 – z1) = 9.81 (0 – 4.3) = – 42.18 J/kg


∴ 1775 × 103 + 300 + (– 42.18) = 2100 × 103 – W
∴ Ws = 324742.18 J/kg
· 3600
Mass flow rate of water = m = 3600 kg/h = 3600 = 1 kg/s

Power is work done per unit time.


·
Power in J/s = Ws in J/kg × m in kg/s
Power developed
∴ by the turbine = 324742.18 J/kg × 1 kg/s
= 324742.18 J/s = 324742.18 W (1 kJ/s = 1 W)
1 hp
Power developed by the turbine = 324742.18 W × 745.7 W = 435.486 ≈ 435.5 hp ... Ans.

Example 2.13 :
Water at 200 kPa and 82°C (355 K), enters a straight pipe with a velocity of 3 m/s, where it
is heated by flue gases from outside. Steam leaves the system at 100 kPa and 150°C (423 K) with
a velocity of 200 m/s. Find the heat that must have been supplied per kg of water flowing.
Data : H for water = 343.3 kJ/kg, H for steam = 2776.3 kJ/kg
Solution :
To obtain Q, we have to make use of the first law equation which is applicable for a steady
flow process.
Basis : 1 kg of water flowing.
For the system under consideration :
z1 = z2 Therefore, ∆ z = 0 ∴ g ∆ z = 0
No work is done.
Therefore, Ws = 0, Q = ?
u1 = 3 m/s ... velocity of water
u2 = 200 m/s ... velocity of steam
H1 = 343.3 kJ/kg = 343.3 × 103 J/kg
H2 = 2776.3 kJ/kg = 2776.3 × 103 J/kg
Chemical Engineering Thermodynamics - I 2.65 First Law of Thermodynamics

Water Steam
z1, u1, H1 u2, H2, z2

Q - flue gases

Fig. E 2.13
The first law of thermodynamics for a steady-state flow process is
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... per unit mass of water

Each term of this equation has the units of J/kg.


∆H = H2 – H1
2 2
∆u2 = u2 – u1
∆z = z2 – z1
1
∴ (2776.3 × 103 – 343.3 × 103) + 2 [(200)2 – (3)2] + 9.81 × 0 = Q – 0

2433 × 103 + 19995.5 + 0 = Q – 0


∴ Q = 2452995.5 J/kg
For 1 kg of water flowing :
Q = 2452995.5 × 1 = 2452995.5 J
= 2452.9955 kJ ≈ 2453 kJ
Heat that must have been supplied per kg of water flowing = 2453 kJ ... Ans.
Example 2.14 :
A system undergoes a process 1-2 in which it absorbs 100 kJ energy as heat and does 40 kJ
work. Then the system follows the process 2-3 during which 30 kJ is rejected as heat while 50 kJ
work is done on it. It is desired to restore the system to the initial state by an adiabatic path.
Calculate (i) the work and heat interactions during the adiabatic process and (ii) the net
work done and the net heat interaction.
Solution :
Process 1-2 : Q1–2 = 100 kJ and W1–2 = 40 kJ ... Heat added to the system and work
done by the system.
The first law for a closed system is
∆U = Q – W
U2 – U1 = Q1–2 – W1–2 = 100 – 40 = 60 kJ
Here Q1–2 and W1–2 are Q and W over the path 1-2.
Chemical Engineering Thermodynamics - I 2.66 First Law of Thermodynamics

Process 2-3 : Q2–3 = – 30 kJ ... Heat rejected by the system


W2–3 = – 50 kJ ... Work is done on the system
U3 – U2 = Q2–3 – W2–3 = – 30 – (– 50) = 20 kJ
U1 – U3 = (U1 – U2) – (U3 – U2)
= – (U2 – U1) – (U3 – U2)
= – 60 – 20 = – 80 kJ
Process 3-1 : U1 – U3 = Q3-1 – W3-1 … first law equation
The process 3-1 by which the system can be restored to the initial state being an adiabatic
process, Q3-1 = 0
– 80 = 0 – W3-1
∴ W3–1 = 80 kJ
For this adiabatic process, Q = 0 and W = 80 kJ ... Ans. (i)
For the cycle 1-2-3-1 :
Net work done, W = W1-2 + W2-3 + W3-1
= 40 – 50 + 80 = 70 kJ ... Ans. (ii)
Net heat interaction, Q = Q1-2 + Q2-3 + Q3–1
= 100 – 30 + 0
= 70 kJ ... Ans. (ii)
[Note that W1-2, i.e., work done can also be denoted by the symbol 1W2.]
Example 2.15 :
A system undergoes a series of processes as a result of which it is finally restored to its
initial state. The work and heat interactions for processes taking place are as given below.
Process ∆U (kJ) Q (kJ) W (kJ)
1-2 – 200 100
2-3 200 – 150 –
3-4 – – – 250
4-1 50 – 300
Calculate the unknown quantities in the above table and the net work and heat interactions.
Solution :
Process 1-2 : Q1-2 = 200 kJ ... Heat added to the system
and W1-2 = 100 kJ ... Work done by the system
Chemical Engineering Thermodynamics - I 2.67 First Law of Thermodynamics

The first law equation for a non-flow process is


∆U = Q – W
U2 – U1 = Q1-2 – W1-2
= 200 – 100 = 100 kJ
∴ ∆U1-2 = 100 kJ ... Ans.
Process 2-3 : U3 – U2 = 200 kJ
and Q2-3 = – 150 kJ ... Heat removed from the system
∆U2-3 = U3 – U2 = Q2-3 – W2-3
200 = – 150 – W2-3
W2-3 = – 350 kJ ... Ans.
Process 3-4 : W3-4 = – 250 kJ
U4 – U3 = (U4 – U1) + (U1 – U2) + (U2 – U3)
= – (U1 – U4) – (U2 – U1) – (U3 – U2)
∆U3-4 = – 50 – 100 – 200 = – 350 kJ ... Ans.
U4 – U3 = – 350 = Q3-4 – W3-4
Q3-4 = – 350 + (– 250) = – 600 kJ ... Ans.
[(U4 – U1) = – (U1 – U4) = – 50 kJ ... See the data for the process 4-1]
Process 4-1 : U1 – U4 = 50 kJ, W4-1 = 300 kJ
U1 – U4 = Q4-1 – W4-1
50 = Q4-1 – 300
Q4–1 = 50 + 300 = 350 kJ ... Ans.
Net work done, W = W1-2 + W2-3 + W3-4 + W4-1
= 100 + (– 350) + (– 250) + 300 = – 200 kJ ... Ans.
Net heat interaction, Q = Q1–2 + Q2-3 + Q3-4 – Q4-1
= 200 + (– 150) + (– 600) + 350
= – 200 kJ … Ans.
Example 2.16 :
A certain ideal gas with Cp/Cv = 1.4 is filled in a vessel of volume 1 m3 till the pressure in it
becomes 5 bar gauge. The temperature is 27°C (300 K).
(i) Calculate the mass of gas in the vessel and the values of Cp and Cv of the gas.
(ii) If 50 kJ of heat is given to the gas in this closed vessel, determine the resulting pressure
and temperature.
Molecular weight of gas = 29 and atmospheric pressure = 100 kPa.
Chemical Engineering Thermodynamics - I 2.68 First Law of Thermodynamics

Solution :
Cp
Cv = γ = 1.4
R
(i) We have, Cv = where, R = 8.31451 kJ/(kmol·K)
γ–1
8.31451
= 1.4 – 1 = 20.786 ≈ 20.79 kJ/(kmol·K) ... Ans. (i)
Cp – Cv = R
∴ Cp = R + Cv
=
8.31451 + 20.79
29.10451 ≈ 29.10 kJ/(kmol·K)
= ... Ans. (i)
P =
5 bar gauge (bar.g) = 500 kPa gauge (kPa.g)
Absolute pressure =
gauge pressure + atmospheric pressure
∴ P =
500 + 100 = 600 kPa
Given : T 300 K, V = 1 m3
=
For an ideal gas, PV =
nRT
PV
∴ n = RT , R = 8.31451 m3·kPa/(kmol·K)
600 × 1
= = 0.2405 kmol
8.31451 × 300
Mass of the gas = Moles × Molecular weight
= 0.2405 kmol × 29 kg/kmol = 6.9745 kg ... Ans. (i)
(ii) Heat added to the gas = Q = 50 kJ
It is a constant volume process ... adding heat to the gas in the closed vessel.
For this constant volume process, Q is given by
Q = nCv (T2 – T1) , T1 = 300 K
50 = 0.2405 × 20.79 (T2 – 300)
∴ T2 = 310 K (37°C) ... Ans. (ii)
P1 P2
At constant volume, T1 = T2

P2 = T2 T  = 310 300 = 620 kPa = 6.20 bar


P1 600
∴ ... Ans. (ii)
 1  
Example 2.17 :
A certain mass of air, initially at a pressure of 480 kPa and temperature 190°C (463 K) is
expanded adiabatically to a pressure of 94 kPa. It is then heated at constant volume until it
attains its initial temperature when the pressure is found to be 150 kPa. State the type of
compression necessary to bring the system back to its original pressure and volume. Find (i) the
index of adiabatic expansion and (ii) the work done per kg of air.
Data : Molecular weight of air = 29.
Chemical Engineering Thermodynamics - I 2.69 First Law of Thermodynamics

Solution :
Temperatures at points 1 and 3 are the same.
Therefore, the process 3-1 is an isothermal process. 1
For process 3-1 :
3
P1V1 = P3V3 as T is constant
P
V3 P1
V1 = P3 2

where P1 = 480 kPa and P3 = 150 kPa


V3 480 V
V1 = 150 = 3.2 Fig. E 2.17
The process 2-3 is a constant volume process.
∴ V2 = V3
V2
∴ V1 = 3.2, since V3 = V2
For the process 1-2 [adiabatic expansion process],
γ γ
P1V1 = P2V2
Taking the logarithm of both sides yields
ln P1 + γ ln V1 = ln P2 + γ ln V2
ln (P2/P1)
∴ γ = ln (V /V )
1 2

where P1 = 480 kPa, P2 = 94 kPa


V2 V1 1
V1 = 3.2 ∴ V2 = 3.2 = 0.3125
ln (94/480)
∴ γ = ln (0.3125) = 1.40178 ≈ 1.402
The index of adiabatic expansion is 1.402 ... Ans. (i)
The work done during an adiabatic expansion process (process 1-2) is given by
P1V1 – P2V2 R
W1-2 = = (T – T2)
γ–1 γ–1 1
For the process 1-2, we have :
T2 P2(γ–1)/γ =  94 (1.4 – 1)/1.4 = 0.6276
T1 = P1 480
∴ T2 = 0.6276 T1
= 0.6276 (463) = 290.57 ≈ 290.6 K
8.31451
∴ W1-2 = 1.4 – 1 (463 – 290.6) = 3583.5 kJ/kmol
Chemical Engineering Thermodynamics - I 2.70 First Law of Thermodynamics

The process 2-3 is a constant volume process.


∴ W2-3 = 0
For the process 3-1 (isothermal process).
The work done in an isothermal process is given by
W3-1 = RT ln (P3/P1)
where T1 = 463 K, P3 = 150 kPa and P1 = 480 kPa
W3-1 = 8.31451 × 463 ln (150/480)
= – 4477.7 kJ/kg
Net work done = W1-2 + W2-3 + W3-1
= 3583.5 + 0 – 4477.7 = – 894.2 kJ/kmol
Mol. wt. of air = 29 kg/kmol
∴ The net work done per kg of air is
– 894.2 kJ/kmol
W = 29 kg/kmol = – 30.83 kJ/kg ... Ans.
The minus (–ve) sign indicates that the work is done on the system.
Example 2.18 :
A gas at a pressure of 1.4 MN/m2 and 360°C (633 K) is expanded adiabatically to a pressure
of 100 kN/m2. The gas is then heated at constant volume until it attains 360°C (633 K) where its
pressure is found to be 220 kN/m2 and finally it is compressed isothermally to the original
pressure of 1.4 MN/m2. Sketch the process on P-V diagram and for 0.5 kmol of gas, calculate the
change in internal energy during the adiabatic expansion. Assume ideal gas behaviour.
Solution :
(i) 1.4 MN/m2 and 633 K → 100 kN/m2 and T2 ... Adiabatic
(ii) 100 kN/m2 and T2 → 220 kN/m2 and 633 K ... Constant V
(iii) 220 kN/m2 and 633 K → 1.4 MN/m2 and 633 K ... Constant T
Let us first find the value of adiabatic index (γ). To find γ, let us find T2 by considering the
process 2-3.
For the process 2-3 ... Constant volume process :
P2V2 P3V3 1
T2 = T3
But V2 = V3 as V is constant 3

P2 P3 P
∴ T2 = T3 2
P2
∴ T2 = T3 · P
3
V
Fig. E 2.18
where T3 = 360°C (633 K), P3 = 220 kN/m2 and P2 = 100 kN/m2
100
∴ T2 = 633 × 220 = 287.727 ≈ 287.73 K (14.73°C)
Chemical Engineering Thermodynamics - I 2.71 First Law of Thermodynamics

Consider the process 1-2. It is an adiabatic expansion process. For this process, we have
T2 P2(γ–1)/γ
= P 
T1  1
γ–1
Let = y
γ
T2 P2y
∴ = P 
T1  1
where T2 = 287.73 K, T1 = 633 K, P1 = 1.4 MN/m2 = 1.4 × 103 kN/m2 and P2 = 100 kN/m2

287.73
=  100 y
633 1.4 × 103
y
∴ 0.4545 = (0.07143)
Taking the logarithm of both sides,
ln (0.4545) = y ln (0.07143)
ln (0.4545) – 0.7885
∴ y = ln (0.07143) = – 2.6390 = 0.29878
γ–1
y = = 0.29878
γ
1
1 – = 0.29878
γ
1
= 1 – 0.29878 = 0.70122
γ
∴ γ = 1.426
The adiabatic index is 1.426
The change in internal energy during an adiabatic process is given by
∆U = – W = – Cv (T1 – T2)
∴ ∆U = nCv (T2 – T1) … for n moles of gas
where T1 = 633 K, T2 = 287.73 K and n = 0.5 kmol
R 8.31451
Cv = = 1.426 – 1 = 19.52 kJ/(kmol·K)
γ–1
∆U = 0.50 × 19.52 (287.73 – 633) = – 3369.83 kJ … Ans.
The change in internal energy is – 3369.83 kJ. … Ans.
Example 2.19 :
A non-flow system executes four different thermodynamic processes in sequence
continuously as given below :
1-2 adiabatic compression, 2-3 isobaric heat addition,
3-4 adiabatic expansion and 4-1 constant volume heat rejection.
The temperatures of the four salient points 1, 2, 3 and 4 are respectively 300 K, 700 K,
1500 K and 600 K. The working substance is 1 kg mass of air.
Take for air : Cp = 29.14 kJ/(kmol·K) and Cv = 20.82 kJ/(kmol·K), Mol. Wt. = 29
Determine the net work done, the net heat transfer and the change in internal energy for the
cycle.
Chemical Engineering Thermodynamics - I 2.72 First Law of Thermodynamics

Solution :
1 (T1 = 300 K) – 2 (T2 = 700 K) ... Adiabatic compression
2 (T2 = 700 K) – 3 (T3 = 1500 K) ... Isobaric heat addition
3 (T3 = 1500 K) – 4 (T4 = 600 K) ... Adiabatic expansion
4 (T4 = 600 K) – 1 (T1 = 300 K) ... Constant volume heat rejection
The cycle 1-2-3-4-1 is shown in Fig. E 2.19.
Cp = 29.14 kJ/(kmol·K)
2 3
Cv = 20.82 kJ/(kmol·K)
Cp 29.14 P 4
Cv = γ = 20.82 = 1.3996 ≈ 1.4
Process 1-2 : Adiabatic compression process. 1
For this adiabatic compression process,
V
Heat transfer, Q1-2 = 0
Fig. E 2.19
… being adiabatic process
R
Work transfer = (T – T2)
γ–1 1
8.31451
= 1.4 – 1 (300 – 700) = – 8314.51 kJ/kmol
Convert this work in kJ with the help of amount of air and its molecular weight.
Amount of air = 1 kg, Molecular weight of air = 29 kg/kmol
W1-2 = 29 kg/kmol × (– 8314.51 kJ/kmol)
1 kg
 
W1-2 = – 286.7 kJ
The first law for a closed system is
∆U = Q – W
∆U = ∆U1-2 = Q1-2 – W1-2
∴ ∆U = –W1-2 … as Q = 0
∆U1-2 = – (– 286.7) = 286.7 kJ
Process 2-3 : Isobaric (P = constant) heat addition process.
For this isobaric heat addition process,
Heat transfer = Q = Q2-3 = nCp (T3 – T2)

= 29 × 29.14 × (1500 – 700)


1
 
= 803.86 kJ
n = number of moles of air = Mass of air 1 kg 
 Molecular weight of air = 29 kg/kmol
Chemical Engineering Thermodynamics - I 2.73 First Law of Thermodynamics

Change in internal energy :


∆U2-3 = nCv (T3 – T2)
where n = 1/29, T3 = 1500 K, T2 = 700 K and Cv = 20.82 kJ/(kmol·K)

∆U2-3 = 29 × 20.82 × (1500 – 700) = 574.34 kJ


1
 
We have, ∆U = Q – W
∴ ∆U2-3 = Q2-3 – W2-3
∴ W2-3 = Q2-3 – ∆U2-3
= 803.86 – 574.34 = 229.52 kJ
Process 3-4 : Adiabatic expansion process.
For this adiabatic expansion process,
Heat transfer, Q3-4 = 0
R
Work transfer, W3-4 = (T – T4) … per unit mole
γ–1 3
8.31451
= 1.4 – 1 (1500 – 600) = 18707.65 kJ/kmol

1 kg
∴ W3-4 for the given amount of air = 29 kg/mol × 18707.65 = 645.1 kJ

W = nR (T – T ) … in kJ
 3-4 γ –1 3 4

∆U3-4 = Q3-4 – W3-4
∆U3-4 = 0 – W3-4
= – W3-4 = – 645.1 kJ
Process 4-1 : Constant volume heat rejection process.
For this constant volume heat rejection process,
Work transfer, W4-1 = 0 ... since V is constant, no displacement work.
Heat transfer, Q4-1 = nCv (T1 – T4), where n = 1/29 … number of moles, kmol
1
Q4-1 = 29 × 20.82 (300 – 600)
= – 215.38 kJ
∆U4-1 = Q4-1 – W4-1
∆U4-1 = Q4-1 – 0
= – 215.31 kJ
Net work transfer, Wnet = W1-2 + W2-3 + W3-4 + W4-1
= – 286.7 + 229.52 + 645.1 + 0
= 587.92 kJ ... Ans.
Chemical Engineering Thermodynamics - I 2.74 First Law of Thermodynamics

Net heat transfer, Qnet = Q1-2 + Q2-3 + Q3-4 + Q4-1


= 0 + 803.86 + 0 + (– 215.38)
= 588.48 kJ ... Ans.
Change in internal energy for the cycle = ∆Unet
∆Unet = ∆U1-2 + ∆U2-3 + ∆U3-4 + ∆U4-1
= 286.7 + 574.34 + (– 645.1) + (– 215.38)
= 0.56 kJ ... Ans.
Check : ∆Unet = Qnet – Wnet
= 588.48 – 587.92 = 0.56 kJ ... Ans.
Example 2.20 :
1 kmol of a gas [Cp = 30 kJ/(kmol·K)] that can be approximated by the ideal gas equation
PV = nRT [R = 8.31451 kJ/(kmol·K)] is initially at 27°C (300 K) and 1 bar (100 kPa). It is then
heated at constant pressure to 127°C (400 K) and compressed isothermally to its initial volume.
Find ∆U, ∆H, Q and W.
Solution :
Given : 1 kmol of gas with Cp = 30 kJ/(kmol·K).
∴ n = 1 kmol
The gas is heated at a constant pressure of 1 bar from 300 K to 400 K and is then compressed
at 400 K to its initial volume.
Hint : Based on the above statement, the actual process can be considered to consist of the
following two steps :
(i) A constant pressure process wherein the gas is heated to the final temperature
T2 (400 K, from T1 = 300 K) and the volume increased to V2 (from the initial
volume V1).
(ii) A constant temperature process in which the gas is compressed to the final pressure P1
and the final volume V1 (= initial volume, from V2).
Calculate separately the changes in U, Q, W and H for both the steps mentioned above and
then sum them to get the answers.
Step (i) : Temperature is raised from 300 to 400 K at a constant pressure of 1 bar (100 kPa).
For a constant pressure process :
∆H = nCp ∆T and Q = ∆H
where n = 1 kmol, Cp = 30 kJ/(kmol·K) and ∆T = T2 – T1 = 100 K, where T1 = 300 K
and T2 = 400 K
∴ ∆H = 1 × 30 (400 – 300)
= 3000 kJ
Q = ∆H = 3000 kJ
Let us find V1 (initial volume).
Chemical Engineering Thermodynamics - I 2.75 First Law of Thermodynamics

P1 = 1 bar = 100 kPa, T1 = 300 K and R = 8.31451 m3·kPa/(kmol·K)


P1V1 = nRT1
nRT1 1 × 8.31451 × 300
V1 = P1 = 100 = 24.94 m3

Let us find V2 (final volume in this constant pressure heating process).


nRT2
V2 = P2
But, P2 = P1 as P is constant.
nRT2 1 × 8.31451 × 400
V1 = P1 = 100 = 33.26 m3

We have, ∆H = ∆U + P ∆V
∴ ∆U = ∆H – P ∆V, ∆V = V2 – V1
= 3000 – 100 (33.26 – 24.94)
= 2168 kJ
The first law equation for a closed system is
∆U = Q – W
W = Q – ∆U
= 3000 – 2168 = 832 kJ ... Ans.
Step (ii) : The gas is compressed at T = 400 K to its initial volume, i.e., to a volume of
24.94 m3 from a volume of 33.26 m.3.
nRT2
Final pressure reached, P2 = V
2
where T2 = 400 K (as T = constant)
V2 = 24.94 m3
nRT2 1 × 8.31451 × 400
∴ P2 = V = 24.94
2

= 133.35 kPa (1.3335 bar) ... Ans.


So the pressure increases from 100 kPa (P1) to 133.35 kPa (P2).
Internal energy and enthalpy of an ideal gas depend on temperature. Therefore, in an
isothermal process (process with T = constant) they remain constant.
∴ In an isothermal process, ∆U = 0 and ∆H = 0
H = U + PV
∆H = ∆U + ∆ (PV) = ∆U + ∆ (nRT) = 0 + 0 = 0
W for an isothermal process is given by

W = nRT ln P 
P1
 2
where T = 400 K, P1 = 100 kPa, P2 = 133.35 kPa and R = 8.31451 kJ/(kmol.K)
Chemical Engineering Thermodynamics - I 2.76 First Law of Thermodynamics

W = 1 × 8.31451 × 400 ln 133.35


100

 
= – 957.189 ≈ – 957.2 kJ
We have, ∆U = Q – W and ∆U = 0
∴ Q = W
∴ Q = – 957.2 kJ
Now, we will calculate the changes in U, H, Q and W for the given process.
For the actual process, the changes in U, H, Q and W are obtained as the sum of the
respective values in the above two steps.
Quantity Step (i) Step (ii) Actual process
∆U 2168 kJ 0 2168 + 0 = 2168 kJ
∆H 3000 kJ 0 3000 + 0 = 3000 kJ
Q 3000 kJ – 957.2 kJ 3000 + (– 957.2) = 2042.8 kJ
W 832 kJ – 957.2 kJ 832 + (– 957.2) = – 125.2 kJ
∴ For the actual process,
∆U = 2168 kJ, ∆H = 3000 kJ, Q = 2042.8 kJ, and W = – 125.2 kJ ... Ans.
Example 2.21 :
1 kmol of nitrogen gas is undergoing a change of state from the initial condition of
10 bar (1000 kPa) and 7°C (280 K) to the final condition of 1 bar (100 kPa) and 67°C (340 K).
Calculate the changes in internal energy and enthalpy.
Given : (i) Assume ideal gas bahaviour, (ii) Cv = 20.8 kJ(kmol·K) and Cp = 29.1 kJ/(kmol·K).
Solution :
Amount of N2 gas = 1 kmol
The ideal gas equation is
PV = nRT
n = 1 kmol
Initial conditions : P1 = 10 bar = 1000 kPa and T1 = 280 K
Final conditions : P2 = 1 bar = 100 kPa and T2 = 340 K
Hint : In order to calculate the changes in the state functions U and H, we may assume that
the proposed change in state is occurring along the following two-step process, i.e., the given
process is occurring in the following two steps :
Step (i) : A constant volume process in which the pressure is reduced to the final pressure
P2 = 1 bar and the temperature is reduced to T'.
Chemical Engineering Thermodynamics - I 2.77 First Law of Thermodynamics

Step (ii) : A constant pressure process in which the gas is heated from T' to T2 (final
temperature) and thereby volume is increased to V2 (final volume).
Initial state : 10 bar and 280 K
Final state : 1 bar and 340 K
Pressure reduction from 10 bar to 1 bar - it is a constant volume process whereby T also
reduces from 280 K to T'. Temperature rise from T' to 340 K - it is a constant pressure process in
which the volume of the gas increases to V2.
(i) Constant volume process :
Let P', V' and T' be the pressure, volume and temperature of the gas after this step.
Let us find P' and V'.
Initial conditions : P1 = 10 bar = 1000 kPa and T1 = 280 K
nRT1 1 × 8.31451 × 280
V1 = P = 1000 = 2.328 m3
1

V' = V1 = 2.328 m2 ... being a constant volume process


P' = P2 = Final pressure = 1 bar = 100 kPa
P'V' 100 × 2.328
T' = nR = = 27.999 ≈ 28 K
1 × 8.31451
R = 8.31451 m3 · kPa/(kmol.K)
Pressure changes from P1 (1000 kPa) to P' (100 kPa).
Temperature changes from T1 (280 K) to T' (28 K).
For this constant volume process, the change in internal energy is given by
∆U = nCv ∆T = nCv (T' – T1)
= 1 × 20.8 (28 – 280) = – 5241.6 kJ
The relationship between ∆H and ∆U is
∆H = ∆U + ∆ (PV)
= ∆U + V ∆P
= – 5241.6 + 2.328 (100 – 1000)
= – 7336.8 kJ
This process being a constant volume process, W = 0 (no displacement work)
∆U = Q – W , W=0
∴ Q = ∆U = – 5241.6 kJ
(ii) Constant pressure process : Final conditions : P2 = 100 kPa and T2 = 340 K
nRT2
∴ V2 = P
2

1 × 8.31451 × 340
= 100 = 28.269 m3
Volume changes from V' (2.328 m3) to V2 (28.269 m3).
Temperature changes from T' (28 K) to T2 (340 K).
Chemical Engineering Thermodynamics - I 2.78 First Law of Thermodynamics

For this constant pressure process, the change in enthalpy is given by


∆H = nCp ∆T = nCp (T2 – T')
= 1 × 29.1 (340 – 28)
= 9079.2 kJ
For a constant pressure process,
Q = ∆H
∴ Q = 9079.2 kJ
The relationship between ∆H and ∆U is
∆H = ∆U + ∆ (PV)
∴ ∆U = ∆H – ∆ (PV) = ∆H – P ∆V = ∆H – P (V2 – V')
= 9079.2 – 100 (28.269 – 2.328)
= 6485.1 kJ
The first law equation for a closed system is
∆U = Q – W
∴ W = Q – ∆U
= 9079.2 – 6485.1 = 2594.1 kJ
For the actual process, the changes in U and H are obtained as the sum of the respective
values in the above mentioned two steps.
The change in the internal energy for the actual process is equal to the sum of the internal
energy changes in the above mentioned two steps.
∆U = ∆U for the step-1 + ∆U for the step-2
= – 5241.6 + 6485.1 = 1243.5 kJ ... Ans.
Similarly, ∆H = ∆H for the step-1 + ∆H for the step-2
= – 7336.8 + 9079.2 = 1742.4 kJ ... Ans.
Example 2.22 :
In a small refrigerator, SO2 is circulated as the refrigerant. SO2 gas at a pressure of
5 bar (500 kPa) and a temperature of 67°C (340 K) is to be cooled at a constant volume of
0.142 m3 to 20°C (293 K) as a part of the refrigeration cycle.
Calculate (i) the amount of heat liberated, (ii) the work done by the gas, (iii) the final
pressure after cooling and (iv) the change in enthalpy. Treat SO2 as an ideal gas.
Data : The specific heat in J/(mol·K) of SO2 is given as
Cp = 25.736 + 5.796 × 10–2 T – 3.8112 × 10–5 T2 + 8.612 × 10–9 T3, where T is in K.
Solution :
Hint : Calculate the number of moles (n) using the ideal gas law equation based on the initial
conditions, then calculate ∆U, ∆H, W by using the formulae for a constant volume process.
Chemical Engineering Thermodynamics - I 2.79 First Law of Thermodynamics

SO2 is to be cooled from 340 K to 293 K at a constant volume of 0.142 m3.


SO2 is to be treated as an ideal gas. So for this ideal gas, we can write
P1V1
n = RT where P1 = 5 bar = 500 kPa
1

500 × 0.142
= = 0.0251 kmol
8.31451 × 340
Let the final pressure be P2.
nRT2
P2 = V2
where n = 0.0251 kmol, R = 8.31451 m3·kPa/(kmol·K), T2 = 293 K
and V2 = V1 = 0.142 m3 (since V is constant)
0.0251 × 8.31451 × 293
P2 = 0.142 = 430.61 kPa

= 430.61 kPa (4.3061 bar ≈ 4.31 bar)


∴ The final pressure is 4.31 bar ... Ans. (iii)
As it being a constant volume process, W = 0 (no displacement work)
∴ Work done by the gas on cooling is zero. ... Ans. (ii)
For this constant volume process, the change in enthalpy is given by
T2

∆H = n ⌠
⌡ Cp dT , Cp in J/(mol·K), i.e., kJ/(kmol·K)
T1

293 
= 0.0251  ⌠
⌡ ( 25.736 + 5.796 × 10 –2 T – 3.8112 × 10–5 T2 + 8.612 × 10–9 T3) dT

340 
 5.796 × 10–2
= 0.0251 25.736 (293 – 340) + 2 [(293)2 – (340)2]

3.8112 × 10–5 8.612 × 10–9 
4 – (340)4 
– 3 [ (293) ]
3 – (340)3 +
4 [ (293) ] 
= 0.0251 {– 1209.59 – 862.18 + 197.76 – 50.61}
= 0.0251 (– 1925.62) = – 48.33 kJ = – 48.33 × 103 J ... Ans.
For this constant volume process, the change in internal energy is given by
T2

∆U = n ⌠
⌡ Cv dT
T1

But Cv = Cp – R
Chemical Engineering Thermodynamics - I 2.80 First Law of Thermodynamics

T2

∴ ∆U = n ⌠
⌡ (Cp – R) dT
T1

T2 T2
 T2

= n ⌠

⌡ Cp dT – ⌠
⌡ R dT , where ⌠
⌡ Cp dT = – 1925.62 kJ/kg
T1 T1  T1

... Calculated above


= 0.0251 [– 1925.62 – 8.31451 (T2 – T1)], R = 8.31451 kJ/(kmol.K)
= 0.0251 [– 1925.62 – 8.31451 (293 – 340)]
= 0.0251 [– 1925.62 + 390.78]
= – 38.52 kJ
= – 38.52 × 103 J
For a constant volume process, we have
Q = ∆U = – 38.52 × 103 J
where the minus (–ve) sign indicates that heat is liberated.
∴ Heat liberated = 38.52 × 103 J ... Ans.
Extra : ∆H = ∆U + ∆ (PV)
∆H = ∆U + V(∆P) = ∆U + V (P2 – P1)
= – 38.52 + 0.142 (430.61 – 500) = – 48.37 kJ
Example 2.23 :
Carbon dioxide gas is sold commercially at 60 bar (6000 kPa). The gas is leaking through
the outlet valve slowly, so that its temperature may be assumed constant at the ambient
temperature of 27°C (300 K).
(i) Calculate the work done in the expansion of 10 kg of CO2 gas from 60 bar to
1 bar (100 kPa).
(ii) If the temperature were constant at 17°C (290 K), what would be the work done ?
(iii) Find the changes in enthalpy in both the above cases.
Solution :
(i) 10 kg of CO2 gas is expanding from 6000 kPa (60 bar) to 100 kPa (1 bar) at 300 K. It is
an isothermal process.
For this isothermal process, the work done per mole is given by

W = RT ln P  , kJ/kmol
P1
 2
Chemical Engineering Thermodynamics - I 2.81 First Law of Thermodynamics

where R = 8.31451 kJ/(kmol·K), T = 300 K


P1 = 6000 kPa and P2 = 100 kPa

W = 8.31451 × 300 ln  100 


6000

 
= 10212.74 kJ/kmol
Let us obtain W for the given amount of gas, i.e., W in kJ.
Amount of CO2 gas = 10 kg, Molecular weight of CO2 = 44
10
Moles of CO2 gas = n = 44 = 0.227 kmol

∴ Work done = n × W = 0.227 × 10212.74 = 2318.29 kJ ... Ans. (i)


(ii) 10 kg of CO2 gas is expanding from 60 bar to 1 bar at 290 K.

Work done = nRT ln P 


P1
 2
= 0.227 × 8.31451 × 290 ln  100 
6000
 
= 2241 kJ ... Ans. (ii)
(iii) Since the enthalpy of an ideal gas depends on temperature, in an isothermal process
it remains constant. ∴ ∆H = 0
∴ As the expansion of CO2 occurs isothermally in both the cases, ∆H = 0
The changes in enthalpy for both the above cases are zero. ... Ans. (iii)
Example 2.24 :
An ideal gas is compressed adiabatically from 1.5 bar (150 kPa) and 65°C (338 K) to a
pressure of 9 bar (900 kPa). The process is reversible and γ = 1.23.
Calculate : (i) the work of compression, (ii) the temperature at the end of compression,
(iii) the heat transferred, (iv) the change in internal energy and (v) the change in enthalpy.
Solution :
Given : The process of compression of ideal gas is adiabatic. Therefore, heat transferred (Q)
is zero.
Heat transferred = 0 ... Ans. (iii)
The work done in an adiabatic process is given by
P1V1  P2(γ–1) / γ
W = 1 –    … (1)
γ – 1  P1 
where P1 = 1.5 bar = 1.5 × 105 Pa = 150 kPa
P2 = 9 bar = 9 × 105 Pa = 900 kPa
γ = 1.23
Chemical Engineering Thermodynamics - I 2.82 First Law of Thermodynamics

Initial conditions : P1 = 150 kPa and T1 = 338 K, R = 8.31451 m3.kPa/(kmol.K)


RT1 8.31451 × 338
∴ V1 = P = 150 = 18.73 m3/kmol
1

Substituting values of various parameters in Equation (1), we get


1.5 × 105 × 18.73   9 (1.23 – 1) / 1.23
W = 1–
(1.23 – 1)  1.5 
= 12215217.39 × (– 0.398) = – 4861656.5 J/kmol
W = – 4861.6565 kJ/kmol ≈ – 4861.66 kJ/kmol ... Ans. (i)
Since W is negative, work is done on the system.
The initial and final temperatures are related to the initial and final pressures by
(γ–1) / γ
T2 P2
=  
T1 P1
(γ–1) / γ
P2
T2 = T1 P 
 1
= 338 150
900 (1.23 – 1) / 1.23
 
= 472.52 K
Temperature at the end of compression = 199.52°C (472.52 K) ... Ans. (ii)
For a closed system, the first law equation is
∆U = Q – W , Q = 0 for this adiabatic process
∴ ∆U = – W
= – (– 4861.66)
= 4861.66 kJ/kmol ... Ans. (iv)
Let V2 be the final volume. V2 is given by
P1 1/γ
V2 = V1 P 
 2
= 18.73 900
150 1/1.23
= 4.364 m3/kmol
 
∆H for this process is calculated as :
∆H = ∆U + P2V2 – P1V1 ... in J/kmol
= 4861.66 × 103 + 9 × 105 × 4.364 – 1.5 × 105 × 18.73
= 4872841.0 J/kmol
= 4872.84 kJ/kmol ... Ans.
In the above calcuations, P1 and P2 are expressed in Pa, i.e.,
N/m2 [e.g., P2 = 9 bar = 9 × 105 N/m2]. Therefore, PV is in N·m/kmol, i.e., J/kmol
(N/m2) · (m3/kmol) = N·m + J/kmol . So ∆U is taken in J/kmol and not in kJ/kmol.
 kmol 
Chemical Engineering Thermodynamics - I 2.83 First Law of Thermodynamics

Example 2.25 :
A rigid and insulated tank of volume 2 m3 is divided into two equal compartments by a
partition. One compartment contains an ideal gas at 1 MPa and 327°C (600 K), while the other
compartment contains the same gas at 0.1 MPa and 27°C (300 K). Find the final temperature
and pressure of the gas in the tank if the partition gets punctured.
Take γ = 1.4 for the gas.
Solution :
Volume of tank = V = 2 m3
2
Volume of each compartment = 2 = 1 m3 ... tank is divided into two equal compartments

Compartment A : PA = 1 MPa = 103 kPa, VA = 1 m3 and TA = 600 K


Moles of gas in the compartment A,
PAVA
nA = RT
A

103 × 1
=
8.31451 × 600
= 0.2 kmol
Compartment B : PB = 0.1 MPa = 100 kPa, VB = 1 m3 and TB = 300 K
Moles of gas in the compartment B,
PBVB
nB = RT
B

100 × 1
=
8.31451 × 300
= 0.04 kmol
Total moles of gas in the tank = n = 0.2 + 0.04 = 0.24 kmol
Since there is a spontaneous mixing of the gas from two compartments, no work is required
for the same.
As there is no work done on or by the system, W = 0
Since the tank is insulated, Q = 0
The first law for a closed system is
∆U = Q – W
∴ ∆U = 0 – 0 = 0
U2 – U1 = 0
∴ U2 = U1
Take U = 0 at To = 0 K
Initial internal energy of the gas in A and B = nACvTA + nBCvTB
Let T be the final temperature of the gas after complete mixing.
Chemical Engineering Thermodynamics - I 2.84 First Law of Thermodynamics

Final internal energy of the gas = nCvT


∴ nACvTA + nBCvTB = nCvT
nATA + nBTB 0.20 × 600 + 0.04 × 300
∴ T = n = 0.24 = 550 K
Final temperature of the gas = 550 K (277°C) ... Ans.
Let P be the final pressure of the gas.
nRT 0.24 × 8.31451 × 550
P = V = 2
= 548.75 kPa = 0.5487 MPa ≈ 0.55 MPa ... Ans.
Example 2.26 :
One mole of an ideal gas contained in a piston-cylinder assembly is compressed from
100 kPa and 27°C (300 K) till its volume is reduced to 1/15 of the original volume. The process
of compression is polytropic with n = 1.2. Determine (i) the final temperature and pressure of
the gas, (ii) the work done on the gas, and (iii) the heat interaction.
Solution :
The path followed by the gas during compression is given by
PV1.2 = Constant
1.2 1.2
∴ P1V1 = P2V2


V11.2
P2 = P1 V 
 2
V1 = Original/initial volume
V2 = Final volume
1
= 15 of V1

V2 = 15 V1 , P1 = 100 kPa


1

 


 V1 1.2
P2 = 100 (1/15) V 
 1

P2 = 100 [15]1.2 = 2578 kPa = 25.78 bar ... Ans. (i)


For an ideal gas,
P1V1 P2V2
T1 = T2
P2 V2
∴ T2 = T1 · P · V
1 1

where T1 = 300 K, P2 = 2578 kPa, P1 = 100 kPa, V2 = (1/15) V1


2578 (1/15) V1
T2 = 300 × 100 × V1
= 515.6 K (242.6°C) ... Ans. (i)
Chemical Engineering Thermodynamics - I 2.85 First Law of Thermodynamics

The work done in a polytropic process is given by


P1V1 – P2V2 R (T1 – T2)
W = n–1 = n–1
R (T1 – T2)
= n–1
8.31451 (300 – 515.6)
= 1.2 – 1
= – 8963 J/mol or kJ/kmol
∴ W = – 8.963 kJ/mol ... Ans. (ii)
The minus (–ve) sign indicates that the work is done on the gas.
Work done on the gas = 8.963 kJ/kmol … Ans. (ii)
The change in internal energy per mole in this process is given by
∆U = Cv (T2 – T1)
R
= (T – T1)
γ–1 2
8.31451
= 1.4 – 1 (515.6 – 300) = 4481.5 kJ/kmol = 4.4815 J/mol
The first law for this non-flow process is
∆U = Q – W
∴ Q = ∆U + W
= 4.4815 + (– 8.9630)
= – 4.4815 kJ/mol ... Ans. (iii)
The minus (–ve) sign indicates that the heat is given out by the system.
Heat given out
by the system = Heat interaction with the surroundings
= 4.4815 kJ/mol … Ans. (iii)
nRT1 1 × 8.31451 × 300
Extra : V1 = P = 100 = 24.94 m3
1

V2 = 15 V1 = 15 × 24.94 = 1.663 m3


1 1
   
2 2
dV
Work done : ⌠ P dV = ⌡
W = ⌡ ⌠ C n as PVn = C
V
1 1

C –n+1 –n+1 n
W = 1 – n [V2 – V1 ] and P1V1 = C
n
P1V1 –n+1 –n+1
∴ W = 1 – n [V 2 – V 1 ]
100 (24.94)1.2
= 1 – 1.2 [(1.663)–1.2 + 1 – (24.93)–1.2 + 1]
= – 8961 kJ/kmol or J/mol = – 8.961 kJ/mol … same as above
Chemical Engineering Thermodynamics - I 2.86 First Law of Thermodynamics

Example 2.27 :
An evacuated tank is connected to a pipe carrying steam at 14 bar and 72°C (345 K) through
a valve. The valve is opened and the tank is filled with steam until the pressure in the tank
becomes 14 bar and then the valve is closed. Find the final temperature of the steam if the
process is adiabatic. Assume the kinetic and potential energy changes to be negligible.
Data : For steam at 14 bar and 345 K : H = 3097 kJ/kg
For steam at 14 bar and 673 K : U = 2952.5 kJ/kg
For steam at 14 bar and 773 K : U = 3121.1 kJ/kg
Solution :
For charging of a tank, the first law of thermodynamics gives
m2U2 – m1U1 = (m2 – m1) H + Q … (1)
where m1 = Initial mass of steam in the tank of internal energy U1
m2 = Final mass of steam in the tank of internal energy U2
(m2 – m1) = Mass of steam that entered the tank of enthalpy H
For this case :
m1 = 0 ... as the tank is evacuated/empty
The process is adiabatic, so Q = 0
Substituting for m1 and Q, Equation (1) becomes
∴ m2U2 – 0 = (m2 – 0) H + 0
∴ m2U2 = m2H
∴ U2 = H
For steam at 14 bar and 345 K : H = 3097 kJ/kg ... given for the steam in the pipe
∴ U2 = H = 3097 kJ/kg
Now we have to find out the final temperature of the steam in the tank having
U = 3097 kJ/kg. Let it be T2 K.
P in the tank (final) = 14 bar ... given
For steam at 14 bar and 673 K : U = 2952.5 kJ/kg
For steam at 14 bar and 773 K : U = 3121.1 kJ/kg
Our value of U, i.e., U2 = 3097 kJ/kg is in between the above two values of U and as U is a
function of T only, T is in between 673 K and 773 K.
∴ T is in between 673 K and 773 K as U2 = 3097 kJ/kg
∆T = 773 – 673 = 100 K
and ∆U = 3121.1 – 2952.5 = 168.6 kJ/kg
∆U
∴ U2 = U at 673 K + · ∆T' where ∆T' = (T2 – 673)
∆T
Chemical Engineering Thermodynamics - I 2.87 First Law of Thermodynamics

168.6
3097 = 2952.5 + 100 · ∆T'

∴ ∆T' = 85.75 K
∆T' = T2 – 673 = 85.75
∴ T2 = 673 + 85.75 = 758.75 K ≈ 759 K
Final temperature of steam = 759 K ... Ans.
Example 2.28 :
A rigid insulated tank of volume 4 m3 is divided into two equal parts by a membrane. On one
side of the membrane, there is a gas A at 5 bar and 77°C (350 K) and on the other side, there is a
gas B at 10 bar and 177°C (450 K). Both A and B are ideal gases with Cv values (5/2) R and
(7/2) R respectively. The membrane is suddenly ruptured to allow the gases to mix. Calculate the
final temperature and pressure that would be reached (after gases get mixed).
Solution :
Volume of tank = V = 4 m3
V 4
Volume of each compartment of the tank = 2 = 2 = 2 m3 ... as the tank is divided into
two equal parts.
Before mixing :
Compartment 1 : Gas A, V = 2 m3, T = 350 K and P = 5 bar (500 kPa)
PV
Moles of gas A = nA = RT
500 × 2
=
8.31451 × 350
= 0.3436 kmol
Compartment 2 : Gas B, V = 2 m3,
T = 450 K and P = 10 bar = 1000 kPa
PV
Moles of gas B = nB = RT
1000 × 2
=
8.31451 × 450
= 0.5345 kmol
 Total moles of gas  = Moles of A + Moles of B
mixture after mixing
= 0.3436 + 0.5345
= 0.8781 kmol
Volume after mixing = Volume of tank = 4 m3
= CV of gas A = 2 R = 2.5 R kJ/(kmol·K)
5
CV, A
 
= CV of gas B = 2 R = 3.5 R kJ/(kmol·K)
7
CV, B
 
Chemical Engineering Thermodynamics - I 2.88 First Law of Thermodynamics

After mixing :
Mole fraction of A = 0.3436
 in the gas mixture  (0.3436 + 0.5345)
0.3436
xA = 0.8781 = 0.3913

Mole fraction of B = 0.5345


 in the gas mixture  0.8781
xB = 0.6087
CV, m = CV of the gas mixture = xA CV, A + xB CV, B
= 0.3913 (2.5 R) + 0.6087 (3.5 R)
= 3.1087 R kJ/(kmol·K)
For this mixing process of gases A and B,
W = 0 ... no work is done on or by the system
As this mixing process is carried out in the insulated tank,
Q = 0 ... no heat transfer to or from the system
The first law equation for a closed system is
∆U = Q – W
∆U = 0 – 0
∆U = 0
∴ U1 = U2
U1 = Initial internal energy (before mixing of A and B)
U2 = Final internal energy of the gas mixture containing A and B
Take U = 0 at To (base temperature) = 0 K
Let T be the final temperature of the gas mixture.
We have, U1 = U2
∴ nA CV, A TA + nB CV, B TB = nm CV, m T … taking U = 0 at T = 0 K
where nA = 0.3436 kmol, CV, A = 2.5 R kJ/(kmol·K), TA = 350 K ... temperature of gas A
nB = 0.5345 kmol, CV, B = 3.5 R kJ/(kmol·K), TB = 450 K
nm = 0.8781 kmol, CV, m = 3.1087 R kJ/(kmol·K),
T = final temperature of the gas mixture
∴ 0.3436 × 2.5 R × 350 + 0.5345 × 3.5 R × 450 = 0.8781 × 3.1087 R × T
∴ T = 418.53 K (145.53°C)
Final temperature of the gas mixture = 145.53°C (418.53 K) ... Ans.
Let P be the final pressure (after mixing of the gases A and B).
nmRT
P = V
Chemical Engineering Thermodynamics - I 2.89 First Law of Thermodynamics

where nm = 0.8781 kmol, V = Volume of tank = 4 m3 , T = 418.53 K


and R = 8.31451 m3.kPa/(kmol.K)
0.8781 × 8.31451 × 418.53
P = 4
= 763.9 kPa = 7.639 bar
Final pressure = 7.639 bar (763.9 kPa) ... Ans.
Example 2.29 :
Two vessels, both containing N2 gas, are connected by a valve. The valve is opened to allow
the contents to mix and achieve an equilibrium temperature of 27°C (300 K).
The following information is known before mixing the contents of the two vessels :
Vessel A : P = 1.5 MPa, T = 50°C (323 K) and amount of N2 = 0.5 kmol
Vessel B : P = 0.6 MPa, T = 20°C (293 K) and amount of N2 = 2.5 kg
(i) Calculate the final equilibrium pressure and the heat transferred to the surroundings.
(ii) If the vessels had been perfectly insulated, what would have been the final temperature
and pressure ?
Cp
Take C = 1.4 and assume ideal gas behaviour.
v
Solution :
(i) To find the final equilibrium pressure and the heat transferred to the surroundings - after
mixing the contents of the two vessels. Both the vessels contain N2 gas.
Vessel A : PA = 1.5 MPa = 1500 kPa, TA = 323 K and nA = 0.5 kmol
Find V of the vessel A.
nARTA
VA = PA
0.50 × 8.31451 × 323
= 1500
= 0.895 m3
Vessel B : PB = 0.60 MPa = 600 kPa, TB = 293 K
Amount of N2 gas = 2.5 kg
Molecular weight of N2 = 28
2.5
Moles of N2, nB = 28 = 0.08928 kmol
nBRTB
VB = P
B

0.08928 × 8.31451 × 293


= 600
= 0.362499 ≈ 0.3625 m3
Chemical Engineering Thermodynamics - I 2.90 First Law of Thermodynamics

After mixing :
Total volume, V = Volume of vessel A + Volume of vessel B
= 0.8950 + 0.3625 = 1.2575 m3
Total moles of N2 gas, n = nA + nB
n = 0.50 + 0.08928 = 0.58928 kmol
Final temperature, T = 300 K (given)
Let P be the final pressure of the contents after mixing.
PV = nRT
nRT
P = V

0.58928 × 8.31451 × 300


= 1.2575
= 1168.88 ≈ 1168.9 kPa ≈ 1.169 MPa ... Ans. (i)
Work transfer is zero, i.e., no work is done on or by the system as it being a gas mixing
process (spontaneous).
∴ W = 0
The first law equation for this mixing process is
∆U = Q – W
∴ ∆U = Q as W = 0
∴ Heat transferred = Change in internal energy = ∆U = U2 – U1
Take U = 0 at To = 0 K … base temperature
Let us find Cv of N2.
Cp
Given : C = 1.4 = γ
v

We know, Cp – Cv = R
Cv = Cp – R
Cp R
1 = C – C
v v

R Cp
Cv = Cv – 1
R R
∴ Cv = C /C – 1 =
p v γ–1
8.31451
Cv = 1.4 – 1 = 20.786 ≈ 20.79 kJ/(kmol·K)
Chemical Engineering Thermodynamics - I 2.91 First Law of Thermodynamics

Initial internal energy, U1 = nA CV, A TA + nB CV, B TB


= 0.50 × 20.79 × 323 + 0.08928 × 20.79 × 293
= 3901.43 kJ
[U1 = U of A + U of B before mixing]
We have, n = 0.58928 kmol and T = 300 K
Final internal energy (after mixing),
U2 = nCvT
= 0.58928 × 20.79 × 300
U2 = 3675.34 kJ
∴ Heat transferred, Q = U2 – U1
= 3675.34 – 3901.43
= – 226.09 kJ ... Ans. (i)
The negative sign indicates that heat is given out by the system.
or Heat transferred to the surroundings = 226.09 kJ … Ans. (i)
(ii) To find P and T in case of insulated vessels :
As both the vessels are insulated, Q = 0
No work is done on or by the system during this gas mixing process.
∴ W = 0 … as it being spontaneous
The first law equation application here is ∆U = Q – W.
∆U = Q – W = 0 – 0 = 0
∆U = 0 and U2 – U1 = ∆U
∴ U1 = U2
Initial internal energy = Final internal energy
 (i.e.‚ before mixing)   (i.e.‚ after mixing) 
nA CV, A TA + nB CV, B TB = nCV T ... Taking U = 0 at To = 0 K
where T is the final temperature
0.50 × 20.79 × 323 + 0.08929 × 20.79 × 293 = 0.58928 × 20.79 × T
∴ T = 318.45 K (45.45°C) ... Ans. (ii)
Let P be the final pressure attained.
nRT
P = V
where n = 0.58928 kmol, T = 318.45 K and V = Total volume = 1.2575 m3
R = 8.31451 m3·kPa/(kmol·K)
0.58928 × 8.31451 × 318.45
∴ P = 1.2575
= 1240.77 ≈ 1240.8 kPa
≈ 12.41 bar)
Final pressure attained = 1240.8 kPa (≈ ... Ans. (ii)
Chemical Engineering Thermodynamics - I 2.92 First Law of Thermodynamics

Example 2.30 :
Methane is stored in a tank of volume 5.7 × 10–2 m3 at a pressure of 15 bar and
21°C (294 K). The tank valve is partially opened allowing the gas to flow to a gas holder,
where pressure is held constant at 1.15 bar. Determine the mass of methane removed when the
pressure in the tank dropped to 5 bar under the following conditions :
(i) If the process occurred slowly so that the temperature was constant.
(ii) If the process occurred so rapidly that the heat transferred was negligible. Assume ideal
gas behaviour and take γ = 1.4.
Solution :
Volume of the tank = V = 5.7 × 10–2 m3
The gas is at Pi = 15 bar = 1500 kPa and Ti = 294 K … initial P and T
Moles of methane gas (initially, i.e., before discharging) = ni
PiVi
ni = RT
i
where R = 8.31451 m3 · kPa/(kmol·K) and Vi = V
1500 × 5.7 × 10–2
∴ ni =
8.31451 × 294
= 0.03498 kmol
(i) The process of discharging of methane gas occurs so slowly that the tank exchanges
(receives) heat with (from) the surroundings such that the temperature remains constant. Energy
of the gas in the tank thus remains constant.
Final pressure, Pf = 5 bar = 500 kPa
Final temperature, Tf = Ti = 294 K ... (since T is constant)
Final volume, Vf = Vi = V = 5.7 × 10–2 m3
Final moles of methane gas in the tank,
PfVf
nf = RT
f
500 × 5.7 × 10–2
=
8.31451 × 294
= 0.01166 kmol
Moles of methane gas removed from the tank
= ni – nf
= 0.03498 – 0.01166
= 0.02332 kmol
Molecular weight of methane (CH4) = 16 kg/kmol
Mass of methane removed = 0.02332 × 16 = 0.3731 kg ... Ans. (i)
Chemical Engineering Thermodynamics - I 2.93 First Law of Thermodynamics

(ii) The process of discharging of methane gas occurs so rapidly that the heat transfer is
negligible. Here the gas is suddenly discharged so that there is no enough time for the tank to
exchange heat energy with the surroundings.
As the heat transfer is negligible, Q = 0. So this process is the same as a process of
discharging of gas from the insulated tank.
When an ideal gas discharges from an insulated tank, the temperature of the gas remaining in
the tank is identical with that estimated by using the relation for a reversible adiabatic expansion
process.
So, temperature of the gas remaining in the tank can be estimated using the reversible
adiabatic expansion relation.
For an adiabatic process,
(γ–1) / γ
P2
T2 = T1 P 
 1
Pf(γ–1) / γ
∴ T f = T i P 
 i
where Ti = 294 K, Pi = 1500 kPa, Pf = 500 kPa and γ = 1.4

Tf = 294 1500
500 (1.4 – 1) / 1.4

 
= 214.79 K ≈ 214.8 K
Final moles of methane gas in the tank = nf
PfVf
nf = RT
f
where Pf = 500 kPa, Tf = 214.8 K, Vf = Vi = 5.7 × 10–2 m3
500 × 5.7 × 10–2
nf =
8.31451 × 214.8
= 0.0159578 ≈ 0.01596 kmol
Moles of methane gas removed = (initial – final) moles
= 0.03498 – 0.01596
= 0.01902 kmol
Mass of methane gas removed = 0.01902 kmol × 16 kg/kmol
= 0.3043 kg ... Ans. (ii)
Extra : For discharging of a tank, the first law of thermodynamics gives
niUi + Q = nfUf + (ni – nf) H
Given : Negligible heat transfer, so Q = 0
niUi – nfUf = (ni – nf) H
Chemical Engineering Thermodynamics - I 2.94 First Law of Thermodynamics

We have, U = CvT and H = CpT … taking U = 0 at T = 0 K


niCvTi – nfCvTf = (ni – nf) CpT
Ti + Tf
T can be approximated as 2 .
Ti + Tf
niCvTi – nfCvTf = (ni – nf) Cp  2 
 
PiVi PfVf
ni = RT and nf = RT
i f
But V = Vi = Vf
∴ ni = PiV/RTi and nf = PfV/RTf
PiV PfV PiV PfV Ti + Tf
∴ C T – C T =  –  C  2 
RTi v i RTf v f
RTi RTf p  
Pi Pf Ti + Tf
PiCv – PfCv = T – T  Cp  2 
 i f  
Cp Pi Pf Ti + Tf
Pi – Pf = C T – T   2 
v i f  
Pi Pf Ti + Tf
Pi – Pf = γ T – T   2 
 i f  
where Pi = 15 bar, Pf = 5 bar, Ti = 294 K and γ = 1.4
15 5 294 + Tf
15 – 5 = 1.4 294 – T   2 
 f  
5 294 + Tf
10 = 1.4 0.051 – T   2 
 f  
Obtain Tf by a trial and error procedure. Tf has to be less than 294 K. Assume Tf, calculate
R.H.S. and check for L.H.S. ≈ R.H.S.
For Tf = 250 K, R.H.S. = 11.8
For Tf = 225 K, R.H.S. = 10.455
For Tf = 220 K, R.H.S. = 10.1725
For Tf = 215 K, R.H.S. = 9.88
For Tf = 217 K, R.H.S. = 10
∴ Tf = 217 K
Example 2.31 :
A rigid and insulated tank (V = 1 m3) contains an ideal gas with γ = 1.4 at 0.1 MPa and
27°C (300 K). The tank is connected through a valve, at the top of the tank, to a pipeline
carrying the same gas at 3 MPa and 227°C (500 K). The valve at the top of the tank is opened to
flow the gas into the tank. When the pressure inside the tank reaches 3 MPa, the valve is closed.
Find the final temperature of the gas in the tank and the amount of gas that entered in the tank.
Chemical Engineering Thermodynamics - I 2.95 First Law of Thermodynamics

Solution :
Let ni be the initial moles of gas.
Vi = V = Volume of the tank = 1 m3
Ti = 300 K and Pi = 0.1 MPa = 100 kPa
PiVi 100 × 1
ni = RT = = 0.04 kmol
i 8.31451 × 300
Let nf be the final moles of gas in the tank.
nf – ni = Moles of gas flowed into the tank
Let Pf be the final pressure.
Pf = 3 MPa = 3000 kPa
Let Tf be the final temperature of the gas.
Vf = Final volume = Vi ... as the tank is rigid
We have, from the first law :
 Energy flow into 
Final energy of = Initial energy +  Heat  +  the tank with gas 
 gas in the tank   of the gas  transferred
from the pipeline
nfUf = niUi + Q + (nf – ni) H
Since the tank is insulated, Q = 0
∴ nfUf = niUi + (nf – ni) H
We have, U = CvT and H = CpT … with Uo = 0 at To = 0 K
nf CvTf = niCvTi + (nf – ni) CpT
where T is the temperature of gas in the pipe line = 500 K
Cp
nfTf = niTi + C (nf – ni) T
v
PfVf
nf = RT
f
PiVi
ni = RT
i
Cp
and Cv = γ
PfVf PiVi PfVf PiVi
· T f = · T i + γ  RT – RT  T
RTf RTi  f i
Now, V = Vf = Vi ... for a rigid tank
Pf Pi 
∴ Pf = Pi + γ T – T  T
 f i
Chemical Engineering Thermodynamics - I 2.96 First Law of Thermodynamics

Substituting numerical values gives

3000 = 100 + 1.4  T – 300 × 500


3000 100
 f 
1.4 × 3000 × 500 1.4 × 100 × 500
2900 = Tf – 300
21 × 105
Tf = 2900 + 233.33 = 3133.33

∴ Tf = 670.21 K (397.21°C)
Final moles of gas in the tank,
PfVf
nf = RT
f
PfV
= RT , since Vf = V
f
3000 × 1
=
8.31451 × 670.21
= 0.538 kmol
Amount of gas that entered in the tank
= nf – ni
= 0.538 – 0.04
= 0.498 kmol = 498 mol ... Ans.
Example 2.32 :
An insulated tank (A) of 1 m3 volume initially contains air (assume to behave as an ideal gas,
γ = 1.4) at 10 MPa and 27°C (300 K). The tank (B) is connected to another insulated and
evacuated tank of 1 m3 volume through a valve. The valve is opened to allow the air to flow in
the second tank till the pressure in both the tanks equalises and then closed. Determine the
temperature and pressure of the air in each tank immediately after the valve is closed. Assume
that the air remained in tank A has undergone adiabatic expansion.
Solution :
Tank A : V = 1 m3
Tank B : V = 9 m3
Initially, the tank A is at P = 10 MPa (10000 kPa) and T = 300 K
n = Initial moles of air in the tank A
PV
For an ideal gas, n = RT , R = 8.31451 m3·kPa/(kmol·K)
10000 × 1
=
8.31451 × 300
= 4.01 kmol
Initially, the tank A contains air and the tank B is evacuated.
Therefore, initial moles of air = n + 0 = 4.01 kmol
Chemical Engineering Thermodynamics - I 2.97 First Law of Thermodynamics

Finally : After admitting air in the tank B till equalising pressure and then immediately after
closing the valve :
Let n1 be the kmol of air in the tank A, T1 be the temperature of air in it and P1 be the
pressure.
Let n2 be the kmol of air in the tank B, T2 be the temperature of air in it and P2 be the
corresponding pressure.
Given : Pressure in both the tanks is the same.
P = P1 = P2
Final total moles of air = n1 + n2 kmol
Initial moles of air = Final moles of air
n = 4.01 = n1 + n2
n1RT1
For tank A : P = P1 = V1
8.31451 n1T1
= 1 = 8.31451 n1T1

∴ n1T1 = 0.12027 P … (1)


n2RT2 8.31451 × n2T2
For tank B : P = P2 = V2 = 9
= 0.92383 n2T2
∴ n2T2 = 1.0824 P … (2)
[As P1 = P2 = P]
Both the tanks are insulated, so Q = 0
No work is done by or on the system, so W = 0
The first law of thermodynamics is
∆U = Q – W
∴ ∆U = 0 as Q = 0 and W = 0
Uf – Ui = 0
∴ Ui = Uf, we have : U = nCvT
Initial internal energy = Final internal energy (after flow of air in the tank B)
nCvT 0 n1CvT1 n2CvT2
.......... + .......... = .......... + ..........
tank A tank B tank A tank B
∴ nT = n1T1 + n2T2
where n = 4.01 kmol, T = 300 K ... air temperature in the tank A
4.01 × 300 = n1T1 + n2T2
1203 = n1T1 + n2T2
Chemical Engineering Thermodynamics - I 2.98 First Law of Thermodynamics

Substituting for n1T1 and n2T2 in terms of P from Equations (1) and (2) in the above equation
gives
1203 = 0.12027 P + 1.0824 P
∴ P = 1000.27 kPa ≈ 1000 kPa = 1 MPa
∴ P = P1 = P2 = 1 MPa ... as equal/same pressure in both
A and B
Pressure in the tank A = 1000 kPa = 1 MPa
Pressure in the tank B = 1000 kPa = 1 MPa ... Ans.
Given : Air in the tank A undergoes adiabatic expansion.
For an adiabatic expansion of gas, we have
Tf Pf(γ–1) / γ
= P 
Ti  i
In our case, Tf = T1 and Ti = T of the tank A initially = 300 K
Pf = P1 and Pi = P of the air in tank A initially = 10000 kPa
We have, P1 = 1000 kPa (obtained previously)
γ = 1.4 (given)
T1  1000 (1.4 – 1) / 1.4
∴ =
300 10000
T1 = 300 × 0.51796
= 155.385 ≈ 155.4 K ... Ans.
We have, n1T1 = 0.12027 P
n1 × 155.4 = 0.12027 × 1000, as P = P1 = 1000 kPa
∴ n1 = 0.77394 kmol
≈ 0.774 kmol
We have, n = n1 + n2
n2 = n – n1
n2 = 4.01 – 0.774 = 3.236 kmol
We have, n2T2 = 1.0824 P
3.236 T2 = 1.0824 × 1000, as P = P2 = 1000 kPa
∴ T2 = 334.487 ≈ 334.5 K ... Ans.
Pressure of the air in the tank A = 1000 kPa = 1 MPa
Temperature of the air in the tank A = 155.4 K
Pressure of the air in the tank B = 1000 kPa = 1 MPa
Temperature of the air in the tank B = 334.5 K ... Ans.
Chemical Engineering Thermodynamics - I 2.99 First Law of Thermodynamics

Example 2.33 :
A rigid and insulated tank having a volume of 0.3 m3 initially contains saturated steam at
350 kPa. It is connected through a valve to a pipe line carrying superheated steam at 1400 kPa
and (325°C) 598 K. The valve is then opened to flow the steam into the tank. When the steam
pressure inside the tank rises to 1400 kPa, the valve is closed. Determine the mass of steam that
entered the tank and the final state of steam in the tank.
Data :
Saturated steam at 350 kPa : U = 2548.9 kJ/kg, V = 0.5243 m3/kg
Superheated steam at 1400 K and 598 K : H = 3094.95 kJ/kg
Superheated steam at 1400 K and 623 K : U = 2869.2 kJ/kg, V = 0.2003 m3/kg
Superheated steam at 1400 K and 673 K : U = 2952.5 kJ/kg, V = 0.2178 m3/kg
Solution :
The first law of thermodynamics for the charging of tank is given by

Final energy = Initial energy + due to steam flow +  to heat transfer 


Energy added Energy added due

m2U2 = m1U1 + (m2 – m1) H + Q


m2U2 – m1U1 = (m2 – m1) H + Q … (1)
The tank is insulated, so Q = 0 ... (adiabatic charging)
∴ m2U2 – m1U1 = (m2 – m1) H … (2)
m2 = Final mass of steam in the tank
U2 = Final internal energy
m1 = Initial mass of steam in the tank
U1 = Initial internal energy of steam in the tank
H = Enthalpy of steam that entered the tank
m2 – m1 = Mass of steam entered the tank
At 345 kPa : U1 = U at 350 kPa = 2548.9 kJ/kg
V1 = 0.5243 m3/kg ... Specific volume of steam
V = Tank volume = 0.3 m3
V V
V1 = Initial mass of steam = m
1
V (m3)
m1 = V (m3/kg)
1
0.3
= 0.5243 = 0.5722 kg ... Initial mass of steam in the tank
At 1400 kPa and 598 K :
H = 3094.95 kJ/kg
Chemical Engineering Thermodynamics - I 2.100 First Law of Thermodynamics

Substituting the numerical values in Equation (2) yields


m2U2 – 0.5722 × 2548.9 = (m2 – 0.5722) × 3094.95
m2U2 – 1458.48 = 3094.95 m2 – 1770.67
m2 (3094.95 – U2) = 312.19 … (3)
V 0.3
Final specific volume, V2 = m = m
2 2

0.3
∴ m2 = V ... V2 is in m3/kg
2

Substituting for m2 in Equation (3) gives


0.30
V2 (3094.95 – U2) = 312.19
928.485 = 312.19 V2 + 0.3 U2 ... (4)
To find out V2 (the final specific volume of steam) and U2 (the final internal energy of
steam), we have to adopt a trial and error procedure - make use of the data provided.
After charging is complete, T2 (the final temperature of steam) is greater than 598 K.
Assume T2 → Obtain V2 and U2 → Evaluate R.H.S. of Equation (4) and check for :
L.H.S. ≈ R.H.S. If L.H.S. ≠ R.H.S. → Repeat the procedure.
Given : At P = 1400 kPa and T = 623 K : U = 2869.2 kJ/kg and V = 0.2003 m3/kg
At P = 1400 kPa and T = 673 K : U = 2952.5 kJ/kg and V = 0.2178 m3/kg
Guess 1 : T2 = 623 K
At 623 K : U2 = U at 623 K = 2869.2 kJ/kg
and V2 = 0.2003 m3/kg
928.485 = 312.19 V2 + 0.3 U2 … (4)
L.H.S. Equation (4) = 928.485
and R.H.S. Equation (4) = 312.19 × 0.2003 + 0.30 × 2869.2 = 923.29
∴ L.H.S. ≠ R.H.S.
Guess 2 : T2 = 628 K
Here we have to evaluate U and V at T = 628 K from the data provided ... using interpolation
techniques.
U at 623 K = 2869.2 kJ/kg
U at 673 K = 2952.5 kJ/kg

U2 = U at 628 K = U at 623 K + 
U at 673 K – U at 623 K
 673 – 623  × (628 – 623)
= 2869.2 +  673 – 623  × 5
2952.5 – 2869.2
 
= 2877.53 kJ/kg
Chemical Engineering Thermodynamics - I 2.101 First Law of Thermodynamics

V at 623 K = 0.2003 m3/kg


and V at 673 K = 0.2178 m3/kg

V2 = V at 628 K = V at 623 K + 
V at 673 K – V at 623 K
× (628 – 623)

673 – 623 

= 0.2003 + 
0.2178 – 0.2003
×5
 50 
= 0.20205 m3/kg
We have, L.H.S. = 928.485
R.H.S. = 312.19 V2 + 0.3 U2
= 312.19 × 0.20205 + 0.3 × 2877.53
= 926.34
∴ L.H.S. ≠ R.H.S.
Guess 3 : T = 633 K

U2 = U at 633 K = 2869.2 +  673 – 623  × (633 – 623)


2952.5 – 2869.2
 
= 2885.86 kJ/kg

V2 = V at 633 K = 0.2003 +  673 – 623  × (633 – 623)


0.2178 – 0.2003
 
= 0.2038 m3/kg
L.H.S. of Equation (4) = 928.485
R.H.S. of Equation (4) = 312.19 V2 + 0.30 U2
= 312.19 × 0.2038 + 0.30 × 2885.86 = 929.38
∴ L.H.S. ≠ R.H.S.
Guess 4 : T2 = 630 K

U2 = 2869.2 +  673 – 623  × (630 – 623)


2952.5 – 2869.2
 
= 2880.86 kJ/kg

V2 = 0.2003 +  673 – 623  × (630 – 623)


0.2178 – 0.2003
 
= 0.20275 m3/kg
L.H.S. = 928.485
R.H.S. = 312.19 V2 + 0.30 U2
= 312.19 × 0.20275 + 0.30 × 2880.86
= 627.55 m3/kg
L.H.S. ≠ R.H.S.
Chemical Engineering Thermodynamics - I 2.102 First Law of Thermodynamics

Guess 5 : T = 631 K

U2 = 2869.2 +  673 – 623  × (631 – 623)


2952.5 – 2869.2
 
= 2882.53 kJ/kg

V2 = 0.2003 +  673 – 623  × (631 – 623)


0.2178 – 0.2003
 
= 0.2031 m3/kg
R.H.S. of Equation (4) = 312.19 × 0.2031 + 0.30 × 2882.53
= 928.16
L.H.S. = 928.485
∴ L.H.S. (928.485) ≈ R.H.S. (928.16)
Hence, we have U2 = 2882.53 kJ/kg, V1 = 0.2031 m3/kg and T2 = 631 K
Final state of the steam (superheated) in the tank :
P = 1400 kPa and T = 358°C (631 K) ... Ans.
Final specific volume of the steam = V2 = 0.2031 m3/kg
The final mass of the steam in the tank is
V (tank volume) in m3
m2 = V2 in m3/kg
0.30
= 0.2031 = 1.4771 kg

Mass of the steam that = (Final mass – Initial mass) of the steam in the tank
 entered the tank 
= m2 – m1
= 1.4771 – 0.5722 = 0.9049 kg ≈ 0.905 kg ... Ans.
Example 2.34 :
A part of the wet steam from a pipeline carrying steam at 600 kPa is bled into a throttling
calorimeter. The conditions of steam in the throttling calorimeter are P = 100 kPa and
T = 110°C (383 K). Determine the quality of steam in the pipeline.
Data :
Steam at 600 kPa : HL = 670.56 kJ/kg and HV = 2756.8 kJ/kg
Steam at 100 kPa and 110°C : (383 K) : Superheated
Steam at 100 kPa and 373 K : H = 2676.2 kJ/kg
Steam at 100 kPa and 423 K : H = 2776.4 kJ/kg
Solution :
For a throttling calorimeter, we have
H at the exit = H at the inlet
[Since a throttling process is an isenthalpic process.]
Chemical Engineering Thermodynamics - I 2.103 First Law of Thermodynamics

Let x be the quality (dryness fraction) of steam at 600 kPa.

∴  H at the inlet  = xH + (1 – x) H
to the calorimeter V L

= 2756.8 x + (1 – x) × 670.56
= 670.56 + 2086.24 x kJ/kg … (1)
To find H at the outlet of calorimeter by interpolation :
H at 100 kPa and 373 K = 2676.2 kJ/kg
H at 100 kPa and 423 K = 2776.4 kJ/kg
H at the outlet (exit) = H at 100 kPa and 110°C (383 K)

= 2676.2 +  423 – 373  × (383 – 373)


2776.4 – 2676.2
 
= 2696.24 kJ/kg
Substituting the value of H in Equation (1)
2696.24 = 670.56 + 2086.24x
∴ x = 0.97097 ≈ 0.971
Quality of the wet steam = 0.971 (i.e., 97.1 % dry) ... Ans.
[Dryness fraction of the steam = 0.971]
Example 2.35 :
Find the flow rate of steam required to produce 500 kW from an adiabatic turbine with inlet
conditions of 800 kPa and 400°C (673 K) and exit conditions of 10 kPa and x = 0.95 (where x is
the quality of steam).
Data : For steam at 800 kPa and 673 K : H = 3267.1 kJ/kg
For steam at 10 kPa and x = 0.95 : HL = 191.83 kJ/kg and HV = 2584.70 kJ/kg
Solution :
.
Let m be the mass flow rate of the steam.
Neglecting the changes in potential energy and kinetic energy, the first law of thermodynamics
gives
∆H = Q – Ws ... for unit mass of fluid
The turbine is adiabatic, so Q = 0
∴ ∆H = – Ws
∆H = H2 – H1
At inlet : H1 = H at 800 kPa and 673 K = 3267.1 kJ/kg
At outlet : H2 = H at 10 kPa with x = 0.95
= xHV + (1 – x) HL
= 0.95 (2584.70) + (1 – 0.95) × 191.83 = 2465.06 kJ/kg
Chemical Engineering Thermodynamics - I 2.104 First Law of Thermodynamics

∆H = H2 – H1 = 2465.06 – 3267.10 = – 802.04 kJ/kg


∴ – 802.04 = – Ws
∴ Ws = 802.04 kJ/kg
Power output of the turbine is 500 kW = 500 kJ/s
Power is work done per unit time.
.
Power output in kJ/s = Ws in kJ/kg × m in kg/s
.
500 = 802.04 × m
.
∴ m = 0.6234 kg/s
= 2244.24 ≈ 2244 kg/h
The mass flow rate of steam is 2244 kg/h ... Ans.
Example 2.36 :
Nitrogen gas is to be compressed at a rate of 5000 kg/h from 100 kPa and 300 K to 1000 kPa
and 450 K. Cooling water at 300 K enters the compressor at a rate of 7500 kg/h and leaves at
320 K. Determine the power required by the compressor.
Data : Cp for N2 = 1.071 kJ/(kg·K) and Cp for water = 4.1868 kJ/(kg·K)
Solution :
According to the first law of thermodynamics, we have :
∆H = Q – Ws ... Neglecting K.E. and P.E. changes, per unit
mass of fluid
·
∆H = m Cp (T2 – T1) ... For nitrogen gas
· 5000
where m = 5000 kg/h = 3600 = 1.389 kg/s

Cp = 1.071 kJ/(kg·K), T1 = 300 K and T2 = 450 K


∴ ∆H = 1.389 × 1.071 × (450 – 300)
= 223.14 kJ/s = 223.14 kW
Q = Heat added to the system
∴ – Q = Heat removed from the system
Ws = Work done by the system

– Ws = on the system ⇒ to run the compressor


Work done Power required

– Q = Heat removed from the system by cooling water
·
= m Cpw (T2 – T1)
Chemical Engineering Thermodynamics - I 2.105 First Law of Thermodynamics

·
where m = 7500 kg/h = 2.083 kg/s, Cpw = 4.1868 kJ/(kg·K), T2 = 320 K and T1 = 300 K

∴ – Q = 2.083 × 4.1868 × (320 – 300)


∴ – Q = 174.42 kJ/s = 174.42 kW ... rate of heat removal
∆H = Q – Ws
– Ws = ∆H – Q
= 223.14 + 174.42
= 397.56 kW
The power required by the compressor is 397.56 kW
– Ws = 397.56 kW
∴ Ws = – 397.56 kW
where the minus (–ve) sign indicates that the work is done on the system.
∴ Power required by the compressor /Work of compression required = 397.56 kW
... Ans.
Example 2.37 :
A pump is used to transfer a solution (ρ = 1200 kg/m3) from a mixing vessel to a storage tank
at a velocity of 1 m/s. The diameter of a pipe is 80 mm and the difference between the level in the
mixing vessel and the storage tank is 20 m. Both the vessel and tank are open to the atmosphere.
Frictional loss is 300 W.
Determine : (i) the power input to the pump and (ii) the pressure increase over the pump.
Solution :
According to the first law of thermodynamics for a steady flow process, we have
∆u2 ∆P
2 + g ∆z + ρ + Ws + F = 0 ... mechanical energy balance

∆u2 ∆P
– Ws = 2 + g ∆z + +F ... in J/kg ... (1)
ρ
Pipe diameter = 80 mm = 0.08 m
Velocity at discharge = u2 = 1 m/s
Density of the solution = ρ = 1200 kg/m3
Mass flow rate of the solution,
.
m = ρu2A = ρu2 · π/4 · D2
π
= 1200 × 1 × 4 (0.08)2
= 6.0318 kg/s ≈ 6.03 kg/s
Frictional loss, F = 300 W = 300 J/s ... convert it in J/kg
300 J/s
∴ F = 6.03 kg/s = 49.75 J/kg
Chemical Engineering Thermodynamics - I 2.106 First Law of Thermodynamics

P2, u2, z2

u1, P1, z1

Fig. E 2.37
u1 ≈ 0 , P1 = 1 atm, u2 = 1 m/s, P2 = 1 atm
∆P P2 – P1 1–1
= = 1200 = 0 J/kg
ρ ρ
z2 – z1 = 20 m ... level difference (given)
2 2 2 2
∆u2 u2 – u1 u2 –0 u2(1)2
2 = 2 = 2 = 2 = 2 = 0.5 J/kg
u2 ⇒ m2/s2 ⇒ kg·m·m ⇒ kg·m · m ⇒ N·m ⇒ J/kg
 kg·s2 s2 kg kg 
g ∆z = g (z2 – z1) = 9.81 × 20 = 196.2 J/kg
Substituting the numerical values, Equation (1) yields
– Ws = 0.5 + 196.2 + 0 + 49.75 = 246.45 J/kg
Ws = – 246.45 J/kg
where the –ve sign indicates that the work is required to be done on the system.
∴ Work required of by the pump (– Ws) = 246.45 J/kg

.
Power input of the pump = (– Ws) × m

= 246.45 × 6.03
= 1486.0935

≈ 1486 J/s = 1486 W ≈ 1.5 kW ... Ans. (i)


Pressure increase over the pump = (196.2 + 49.75) × 1200
= 295140 N/m2
≈ 2.95 × 105 N/m2 ... Ans. (ii)
Chemical Engineering Thermodynamics - I 2.107 First Law of Thermodynamics

Example 2.38 :
An adiabatic compressor operating under steady-state conditions receives air at 1 bar and
27°C (300 K). The compressor discharges air at 10 bar. Find the power consumption of the
compressor if air flows at a rate of 2 mol/s through it.
Solution :
Air flow rate = 2 mol/s
P1 = 1 bar = 105 N/m2 , T1 = 300 K
P2 = 10 bar = 106 N/m2 , γ = 1.4 for air
n = 2 mol/s = 2 × 10–3 kmol/s
P1 = 105 N/m2 = 100 kPa
R = 8.31451 m3·kPa/(kmol·K)
nRT1
Initial volume, V1 = P … per second
1

2 × 10–3 × 8.31451 × 300


= 100 = 0.0499 m3/s

The work required for the compression of air (Ws) is given by


 P2 (γ–1) / γ
P1V1 1 – P 
γ
Ws = 
γ–1   1 

= 1.4 – 1 × 105 × 0.0499 × 1 – 105


1.4 106 (1.4 – 1) / 1.4
   
= – 16253 J/s ≈ 16.25 kJ/s
= – 16.25 kW
The –ve sign indicates that the work is required, i.e., the work is done on the system for
compression. Since a compressor is a work consuming machine, W always comes as –ve.
∴ –Ws = 16.25 kW
∴ The power consumption of the compressor is 16.25 kW … Ans.
OR
 P2 (γ–1) / γ
RT1 1 – P 
γ
Ws = 
γ–1   1 

= 1.4 – 1 × 8.31451 × 300 × 1 – 105


1.4 106 (1.4 – 1) / 1.4
   
= – 8124.65 kJ/kmol
∴ – Ws = 8124.65 kJ/kmol
Chemical Engineering Thermodynamics - I 2.108 First Law of Thermodynamics

No. of moles of air compressed = 2 mol/s = 2 × 10–3 kmol/s


Now let us obtain the work required in kJ/s.
– Ws = – Ws in kJ/kmol × air flow in kmol/s
∴ – Ws = 8124.05 × 2 × 10–3
= 16.2493 ≈ 16.25 kJ/s = 16.25 kW ... Ans.
Example 2.39 :
Oil at a rate of 1000 kg/min flows from an open reservoir placed at the top of a hill 400 m in
height to another reservoir at the bottom of the hill. Heat at a rate of 1800 kJ/min and work by a
1 hp pump is supplied to the oil on its way to the bottom reservoir. If the mean specific heat of oil
is 3.35 kJ/(kg·K), find the temperature change of the oil.
Solution :
.
Mass flow rate of the oil, m = 1000 kg/min
1000
= 60 = 16.67 kg/s

Heat supplied to the oil, Q = 1800 kJ/min


1800
= 60 = 30 kJ/s

Let us obtain Q in J/kg


30 kJ/s
Q = 16.67 kg/s

= 1.7996 kJ/kg = 1.8 kJ/kg


= 1.8 × 103 J/kg
Work supplied to the oil,
–Ws = 1 hp = 745.7 J/s
745.7 J/s
= 16.67 kg/s = 44.73 J/kg

The first law of thermodynamics for a steady-state flow process gives


1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid

Here, Ws = – 44.73 J/kg ... as work is supplied to the oil


∆u2 ≈ 0, Q = 1.8 × 103 J/kg, z1 = 400 m, z2 = 0
∴ ∆z = z2 – z1 = 0 – 400 = – 400 m
Substituting the values, we get
∆H + 0 + 9.81 (– 400) = 1.8 × 103 – (– 44.73)
∆H = 5.76873 kJ/kg
Chemical Engineering Thermodynamics - I 2.109 First Law of Thermodynamics

We have, ∆H in kJ/kg = Cp ∆T
∆H in kJ/kg
∴ ∆T = C in kJ/(kg·K)
p
5.76873
= 3.35 = 1.722 K
∴ The temperature change of the oil is 1.722 K or 1.722°C ... Ans.
Example 2.40 :
3 kmol N2 gas at 77°C (350 K) is taken in a rigid container weighing 100 kg and is heated to
277°C (550 K). Find the heat that must be supplied.
Data : Cv for N2 gas = 20.8 kJ/(kmol·K)
Specific heat of container = 0.5 kJ/(kg·K)
Solution :
As the container is rigid, V is constant.
For a constant volume process,
∆U = nCv ∆T = nCv (T2 – T1) … for N2 gas
where n = 3 kmol, Cv = 20.8 kJ/(kmol·K), T2 = 550 K and T1 = 350 K
Q for N2 gas = 3 × 20.8 × (550 – 350) = 12480 kJ
Q for container = mC ∆T = mC (T2 – T1)
where m = 100 kg and C = 0.5 kJ/(kg·K)
∴ Q for container = 100 × 0.50 × (550 – 350)
= 10000 kJ
Q that must be supplied to heat the container with its contents is
Q = Q for N2 gas + Q for container
= 12480 + 10000
= 22480 kJ ... Ans.
Extra : 2 kmol of N2 gas is taken in a piston-cylinder arrangement at 500 K. Find the
quantity of heat removed from the gas to cool it to 350 K at constant pressure. Assume that the
heat capacity of the container is negligible.
Cp of N2 gas = 29.1 kJ/(kmol·K)
Solution :
For a constant pressure process,
Q = ∆H = nCp ∆T = nCp (T2 – T1)
where n = 2 kmol, Cp = 29.1 kJ/(kmol·K), T1 = 500 K and T2 = 350 K
∴ Q = 2 × 29.1 × (350 – 500) = – 8730 kJ
–ve sign indicates that heat is removed from the system.
∴ Heat to be removed to cool the gas to 350 K = 8730 kJ ... Ans.
Chemical Engineering Thermodynamics - I 2.110 First Law of Thermodynamics

Example 2.41 :
A steam jet ejector is employed to entrain saturated water vapour at 25 kPa leaving an
evaporator, with the help of high pressure saturated steam at 1000 kPa. It is estimated that
0.75 kg of the vapours from the evaporator are entrained by every one kg of high pressure steam.
If the mixed stream leaving the ejector is at 100 kPa, what will be its temperature ?
Data : H of saturated vapour at 25 kPa = 2618.2 kJ/kg
H of saturated steam at 1000 kPa = 2778.1 kJ/kg
H of superheated steam at 100 kPa and 373.15 K = 2676.2 kJ/kg
H of superheated steam at 100 kPa and 423.15 K = 2776.4 kJ/kg
Solution :
H of vapour = 2618.2 kJ/kg
H of steam = 2778.1 kJ/kg
Given : 1 kg of steam entrains 0.75 kg of vapour
∴ Mixed stream = 1 + 0.75 = 1.75 kg
0.75
Weight or mass fraction of vapour in the mixed stream = xv = 1.75 = 0.4286

1
Weight or mass fraction of steam in the mixed stream = xs = 1.75 = 0.5714

∴ H of mixed stream leaving the ejector is


H = xv Hv + xs Hs = 0.4286 (2618.2) + 0.5714 (2778.1)
= 2709.5668 ≈ 2709.6 kJ/kg
The temperature of superheated steam at 100 kPa and having an enthalpy of 2709.6 kJ/kg
that leaves the ejector can be obtained from the steam tables by interpolation.
H at 100 kPa and 373.15 K = 2676.2 kJ/kg
H at 100 kPa and 423.15 K = 2776.4 kJ/kg
∴ ∆T = 50 K, ∆H = 2776.4 – 2676.2 = 100.2 kJ/kg
∆H 100.2
∴ = 50 = 2.004 kJ/(kg.K)
∆T
Therefore, for 1 K rise in T, ∆H is 2.004 kJ/kg
H of mixed stream = 2709.6
H at 373.15 K = 2676.2
∆H = 2709.6 – 2676.2 = 33.4 kJ/kg
33.4
∴ T = 373.15 + 2.004 = 389.816 ≈ 389.82 K

∴ The temperature of the mixed stream leaving the ejector is 389.82 K ... Ans.
Chemical Engineering Thermodynamics - I 2.111 First Law of Thermodynamics

Example 2.42 :
The work required to compress a gas from 100 kPa and 300 K to 300 kPa pressure is
280 kJ/kg of the gas. The compressed gas is admitted to a nozzle in which its velocity is
increased to 700 m/s. Find the heat removed during compression if the gas enters the
compressor with negligible velocity and leaves the nozzle at 100 kPa and 300 K.
Solution :
Work done per kg of gas compressed (– Ws) = 280 kJ/kg
∴ Ws = – 280 kJ/kg = – 280 × 103 J/kg
i.e., the work is done on the system.
Compressor
P1 = 100 kPa + Nozzle P2 = 100 kPa
T1 = 300 K T2 = 300 K
u1 = 0 u2 = 700 m/s
z1 z2

Fig. E 2.42
As P1 = P2 and T1 = T2
H1 = H2 ∴ ∆H = 0
Assume z1 = z2 ∴ ∆z = z2 – z1 = 0
The first law for a steady flow process gives
1 2 2
(H2 – H1) + 2 (u2 – u1) + g (z2 – z1) = Q – Ws

Each term in this equation has the units of J/kg. Substituting the values, it yields
1
0 + 2 [(700)2 – 0] + 9.81 × 0 = Q – (– 280 × 103)

245000 = Q + 280 × 103


∴ Q = – 35000 J/kg
= – 35 kJ/kg
where the minus (–ve) sign indicates that the heat is removed from the system.
∴ Heat removed during the compression per kg of the gas = 35 kJ ... Ans.
Example 2.43 :
Dry saturated steam at 0.5 MPa enters an adiabatic nozzle at a velocity of 3 m/s. It leaves
the nozzle as dry saturated steam at 0.2 MPa. Find the exit velocity of steam and the cross-
sectional area of the nozzle at the exit if the steam flow rate through the nozzle is 3000 kg/h.
Data : (i) Saturated steam at 0.5 MPa (5 bar) : HV = 2747.5 kJ/kg
(ii) Saturated steam at 0.2 MPa (2 bar) : HV = 2706.3 kJ/kg and
Specific volume = 0.8854 m3/kg
(iii) Assume that the nozzle is horizontal.
Chemical Engineering Thermodynamics - I 2.112 First Law of Thermodynamics

Solution :
. 3600
Steam flow rate, m = 3600 kg/h = 3600 = 1 kg/s

H1, u1, z1 u2, H2, z2

Fig. E 2.43
The nozzle is horizontal and adiabatic.
∴ z1 = z2 and Q = 0
In case of nozzle, there is no shaft work.
∴ Ws = 0
H1 = Enthalpy of dry saturated steam at 5 bar
= 2747.5 kJ/kg = 2747.5 × 103 J/kg
u1 = 3 m/s
H2 = Enthalpy of dry saturated steam at 2 bar
= 2706.3 kJ/kg = 2706.3 × 103 J/kg
u2 = ?, z1 = z2 ∴z2 – z1 = 0
The first law of thermodynamics for a steady-state flow process is
1 2 2
(H2 – H1) + 2 (u2 – u1) + g (z2 – z1) = Q – Ws ... for unit mass of fluid
1 2
(2706.3 × 103 – 2747.5 × 103) + 2 [u2 – (3)2] + 9.81 × 0 = 0 – 0
1 2
– 41200 + 2 u2 – 4.5 + 0 = 0 – 0
1 2
∴ 2 u2 = 41204.5
2
∴ u2 = 2 × 41204.5 = 82409
∴ u2 = (82409)1/2 = 287.069 ≈ 287.07 m/s
∴ Velocity of steam at the exit of the nozzle = 287.07 m/s ... Ans.
At the exit of the nozzle,
Steam flow rate = 1 kg/s
Steam velocity = 287.07 m/s
Specific volume of steam = 0.8854 m3/kg
Cross-sectional area of the nozzle at the exit = A m2
Chemical Engineering Thermodynamics - I 2.113 First Law of Thermodynamics

The mass flow rate and volumetric flow rate are related by
Mass flow rate × Specific volume = Volume flow rate
= Velocity × A
1 × 0.8854 = 287.07 × A
∴ A = 3.084 × 10–3 m2
Exit cross-sectional area of the nozzle = 3.084 × 10–3 m2 ... Ans.
Example 2.44 :
An adiabatic nozzle receives dry saturated steam at 0.6 MPa (6 bar) with negligible velocity.
The steam leaves the nozzle as dry saturated steam at 0.15 MPa (1.5 bar). Calculate the exit
velocity of the steam.
Data : Saturated steam at 6 bar : HV = 2756.80 kJ/kg
Saturated steam at 1.5 bar : HV = 2693.60 kJ/kg
Solution :
Assume that the nozzle is horizontal.
∴ z1 = z2 and ∆z = 0

z1, H1, u1 z2, H2, u2

Fig. E 2.44
u1 = 0 ... given that the steam enters with negligible velocity
H1 = Enthalpy of dry saturated steam at 6 bar
= 2756.80 kJ/kg = 2756.80 × 103 J/kg
z2 = z1 ... horizontal nozzle
H2 = Enthalpy of dry saturated steam at 1.5 bar
= 2693.60 kJ/kg = 2693.60 × 103 J/kg
u2 = Exit velocity of steam = ?
The nozzle is adiabatic. Therefore, Q = 0.
There is no shaft work delivered in a nozzle.
Therefore, Ws = 0
The first law of thermodynamics for a steady-state flow process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid
1 2 2
(H2 – H1) + 2 (u2 – u1) + g (z2 – z1) = Q – Ws
Chemical Engineering Thermodynamics - I 2.114 First Law of Thermodynamics

Substituting the values in the above equation, we get


1 2
(2693.6 × 103 – 2756.8 × 103) + 2 (u2 – 0) + 9.81 (0) = 0 – 0
1 2
∴ 2 u2 = 63200
2
∴ u2 = 2 × 63200
∴ u2 = (2 × 63200)1/2
∴ u2 = 355.53 m/s
Exit velocity of steam = 355.53 m/s ... Ans.
Example 2.45 :
Superheated steam at 4 MPa (40 bar) and 400°C (673 K) enters an adiabatic turbine and
leaves as wet steam of quality 0.90 at 35oC (308 K). The steam enters the turbine with a velocity
of 5 m/s at an elevation of 10 m from the datum and leaves the turbine with a velocity of 20 m/s
at an elevation of 3 m. Calculate the steam consumption rate if the power output of the turbine is
200 MW. Also, calculate the percent error introduced if the changes in K.E. and P.E. are
neglected.
Data : Superheated steam at 4 MPa and 400°C : H = 3215.7 kJ/kg
Wet steam at 35°C : HL = 146.56 kJ/kg and HV = 2565.4 kJ/kg
Solution :
The first law of thermodynamics for a steady-state flow process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid
At inlet/entrance : H1, u1 and z1 H1, u1, z1

u1 = 5 m/s
H1 = H of superheated steam
Ws
H1 = 3215.7 kJ/kg = 3215.7 × 103 J/kg,
z1 = 10 m
At exit : H2, u2 and z2
z2 = 3 m, u2 = 20 m/s
H2, u2, z2
Fig. E 2.45
H2 = H of wet steam of quality (x, dryness fraction) 0.90
= HL (1 – x) + xHV , HL = 146.56 kJ/kg and HV = 2565.4 kJ/kg

= 146.56 (1 – 0.90) + 0.90 × 2565.4


= 2323.516 ≈ 2323.52 kJ/kg = 2323.52 × 103 J/kg
Power output of the turbine = 200 MW = 200 × 103 kW = 200 × 103 kJ/s
Chemical Engineering Thermodynamics - I 2.115 First Law of Thermodynamics

.
Let the steam consumption rate be m kg/s.
Power is work done per unit time.
·
Power in kJ/s = Ws in kJ/kg × m in kg/s
Power in kJ/s
∴ Ws in kJ/kg = ·
m in kg/s
200 × 103 200 × 106
∴ Work done by the turbine = Ws = kJ/kg = J/kg
. .
m m
The turbine is adiabatic. Hence, Q = 0
The first of thermodynamics for a steady state flow process is given by
1 2 2
(H2 – H1) + 2 (u2 – u1) + g (z2 – z1) = Q – Ws

Substituting the values gives


1 200 × 106
(2323.52 × 103 – 3215.7 × 103) + 2 [(20)2 – (5)2] + 9.81 (3 – 10) = 0 –
.
m
.
– 892180 + 187.5 – 68.67 = – 200 × 106 / m
.
892180 – 187.5 + 68.67 = 200 × 106 / m
.
∴ 892061.17 = 200 × 106 / m
. 200 × 106
∴ m = 892061.17 = 224.1998 ≈ 224.2 kg/s
Steam consumption rate = 224.2 kg/s ... Ans.
(i) Neglect K.E. change : We have :
.
– 892180 + 187.5 – 68.17 = – 200 × 106 / m
........
K.E.
Neglecting KE change,
.
– 892180 – 68.67 = – 200 × 106 / m
.
∴ m = 224.15 kg/s
. 224.2 – 224.15
% error introduced in m if K.E. change is neglected = 224.2 × 100 = 0.0223
(ii) Neglect P.E. change :
.
– 892180 + 187.5 – 68.67 = – 200 × 106 / m
..............
P.E.
Chemical Engineering Thermodynamics - I 2.116 First Law of Thermodynamics

Neglecting PE change,
.
– 892180 + 187.5 = – 200 × 106 / m
.
∴ m = 224.217 kg/s
. 224.217 – 224.2
% error introduced in m if P.E. change is neglected = 224.2 × 100 = 0.00758
… Ans.
Example 2.46 :
In a thermal power plant, an adiabatic steam turbine operating under steady-state
conditions receives superheated steam at 3 MPa (30 bar) and 300°C (573 K). The steam enters
the turbine with a velocity of 10 m/s at an elevation of 10 m above the ground and leaves the
turbine as wet steam of quality 0.85 at 45°C (318 K), at an elevation of 4 m above the ground
and with a velocity of 40 m/s. Find the power output of the turbine if the steam flow rate through
the turbine is 3600 kg/h. Also, calculate the percent error introduced if the changes in K.E. and
P.E. are neglected.
Data : Steam at 3 MPa and 300°C : H = 2995.1 kJ/kg
Saturated steam at 45°C : HL = 188.35 kJ/kg and HV = 2583.3 kJ/kg
Solution :
. H1, u1, z1
Steam flow rate = m = 3600 kg/h = 1 kg/s
x - Quality of steam at the exit of the turbine = 0.85
H1 = Enthalpy at the entrance Ws
= Enthalpy of superheated steam
= 2995.1 kJ/kg = 2995.1 × 103 J/kg
u1 = 10 m/s
z1 = 10 m
H2, u2, z2
H2 = Enthalpy at the exit of the turbine
Fig. E 2.46
= Enthalpy of saturated steam
= (1 – x) HL + xHV
where x = 0.85, HL = 188.35 kJ/kg and HV = 2583.3 kJ/kg
∴ H2 = (1 – 0.85) × 188.35 + 0.85 × 2583.3
= 2224.0575 ≈ 2224.06 kJ/kg = 2224.06 × 103 J/kg
u2 = 40 m/s
z2 = 4 m
The first law of thermodynamics for a steady-state process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid
1 2 2
(H2 – H1) + 2 (u2 – u1) + g (z2 – z1) = Q – Ws
Chemical Engineering Thermodynamics - I 2.117 First Law of Thermodynamics

The turbine is adiabatic. Therefore, Q = 0


1
(2224.06 × 103 – 2995.1 × 103) + 2 [(40)2 – (10)2] + 9.81 (4 – 10) = 0 – W

– 771040 + 750 – 58.86 = – Ws


∴ Ws = 770348.86 J/kg ≈ 770349 J/kg
Work done by the turbine = Ws = 770349 J/kg
Ws is always +ve for a turbine, since it is a work producing machine.
.
Mass flow rate of steam = m = 1 kg/s
.
Power is the work done per unit time [P = Ws in J/kg × m in kg/s]
∴ Power output of the turbine = 770349 J/kg × 1 kg/s
= 770349 J/s
= 770349 W
= 770.349 kW ≈ 770.35 kW ... Ans.
770349
= 745.7 [745.7 J/s = 745.7 W = 1 hp]
= 1033.05 ≈ 1033 hp ... Ans.
(i) Neglect change in K.E. :
We have, – 771040 + 750 – 58.86 = – Ws
......
K.E.
Neglecting the K.E. change,
– 771040 – 58.86 = – Ws
∴ Ws = 771098.86 J/kg
.
Power output of the turbine = Ws in J/kg × m in kg/s = 771098.86 × 1
= 771098.86 J/s = 771.098 kW ≈ 771.10 kW
771.1 – 770.35
% error introduced in the power output if K.E. change is neglected = 770.35 × 100
= 0.09735 ≈ 0.097 ... Ans.
(ii) Neglect change in P.E. :
– 771040 + 750 – 58.86 = – Ws
.......
P.E.
Neglecting the P.E. change,
∴ – 771040 + 750 = – Ws
∴ Ws = 770290 J/kg
.
∴ Power output = Ws in J/kg × m in kg/s
= 770290 × 1 = 770290 J/s = 770.29 kW
Chemical Engineering Thermodynamics - I 2.118 First Law of Thermodynamics

% error introduced in the power output if P.E. change is neglected


770.35 – 770.29
= 770.35 × 100 = 0.00779 ... Ans.

% error introduced if both K.E. and P.E. changes are neglected


= 0.097 + 0.00779 = 0.10479 ≈ 0.1048
Hence, this shows that the changes in K.E. and P.E. terms are not significant in the
calculation of the power output of a turbine.
Example 2.47 :
A steam turbine is designed to generate a power output of 9 MW for a mass flow rate of
17 kg/s. The inlet conditions are : P = 3 MPa, T = 450°C (723 K) and u = 200 m/s; the outlet
conditions are : P = 0.5 MPa (saturated vapour) and u = 80 m/s. Determine the heat transfer for
this turbine.
Data : For steam at 3 MPa and 450°C : H = 3343.1 kJ/kg
For steam at 0.5 MPa (saturated vapour) : H = 2746.6 kJ/kg
Solution :
Assuming the potential energy change to be negligible, the first law of thermodynamics for a
steady-state flow process gives
1
∆H + 2 ∆u2 = Q – Ws ... for unit mass of fluid … (1)

At P1 = 3 MPa and T1= 723 K :


H1 = 3343.1 kJ/kg
u1 = 200 m/s
At P2 = 0.5 MPa (saturated vapour) :
H2 = 2746.6 kJ/kg
u2 = 80 m/s
∴ ∆H = H2 – H1 = 2746.6 – 3343.1 = – 596.5 kJ/kg
.
Mass flow rate of steam = m = 17 kg/s
Turbine power output = 9 MW = 9000 kW = 9000 kJ/s
Power output 9000
∴ Work done, Ws = Mass flow rate = 17 = 529.4 kJ/kg

[Power is work done per unit time.


·
Power in kJ/s = Ws in kJ/kg × m in kg/s]
1 2 2 1 2 2
2 (u2 – u1) = 2 [(80) – (200) ]
= – 16800 J/kg = – 16.8 kJ/kg
Chemical Engineering Thermodynamics - I 2.119 First Law of Thermodynamics

Substituting the values in Equation (1), we get


– 596.5 + (– 16.8) = Q – 529.4
∴ Q = – 83.9 kJ/kg
·
Q in kJ/s = Q in kJ/kg × m in kg/s
Q = – 83.9 × 17
= – 1426.3 kJ/s
= – 1426.3 kW

Fig. E 2.47
where the minus (–ve) value of Q indicates that the heat is transferred from the turbine to the
surroundings. This is expected as the steam temperature is higher than that of the surroundings.
Rate of heat transfer from the turbine = 1426.3 kJ/s = 1426.3 kW … Ans.
Example 2.48 :
A steam turbine is designed to operate at a mass flow rate of 1.5 kg/s. The inlet conditions
are : 2 MPa, 400oC (673 K) and 60 m/s; the outlet conditions are : 0.1 MPa, 0.98 (steam quality)
and 150 m/s. The change in elevation from inlet to outlet is 1 m. Calculate the power output of
the turbine if the heat loss is 50 kW. Also, calculate the power output if the changes in kinetic
and potential energy are neglected.
Data : For steam at 2 MPa and 400°C : H = HV = 3467.6 kJ/kg
For steam at 0.10 MPa : HL = 640.23 kJ/kg, HV = 2748.70 kJ/kg
Solution :
At outlet : Quality of steam = x = 0.98
H2 = H of steam at the outlet
= (1 – x) HL + xHV
= (1 – 0.98) × 640.23 + 0.98 × 2748.70
= 2706.53 kJ/kg
u2 = 150 m/s
z2 = 1 m considering z1 = 0 as the change in elevation, i.e., ∆z is 1 m (given)
At inlet : Steam is superheated at 2 MPa and 673 K
H1 = HV at 2 MPa and 673 K = 3467.6 kJ/kg
u1 = 60 m/s
z1 = 0 as the inlet is considered as datum
The first law of thermodynamics for a steady-state flow process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws … per unit mass of fluid … (1)

∆H = H2 – H1 = 2706.53 – 3467.6 = – 761.07 kJ/kg


Chemical Engineering Thermodynamics - I 2.120 First Law of Thermodynamics

1 2 1 1
2 ∆u = 2 [(u2) – (u1) ] = 2 [(150) – (60) ]
2 2 2 2

= 9450 J/kg = 9.45 kJ/kg


Q = Rate of heat transfer
In our case, Q is to be used in kJ/kg.
Given : Heat loss from the turbine = 50 kW = 50 kJ/s.
∴ Q = – 50 kJ/s
·
Mass flow rate = m = 1.5 kg/s
Q in kJ/s
Q in kJ/kg = ·
m in kg/s
– 50 kJ/s
∴ Q = 1.5 kg/s = – 33.33 kJ/kg

The –ve sign is assigned to the value of Q as the heat is lost from the turbine (system) to the
surroundings (it is given in the statement of the problem).
g ∆z = g (z2 – z1) = 9.81 (1 – 0) = 9.81 J/kg = 9.81 × 10–3 kJ/kg
∴ Putting the values obtained in Equation (1), we get
– 761.07 + 9.45 + 9.81 × 10–3 = – 33.33 – Ws
∴ Ws = 718.3 kJ/kg ... power output per kg
Let us find power output in terms of kW.
Power is work done per unit time. Therefore,
·
Power, kJ/s = Ws, kJ/kg × m , kg/s
Power output of the turbine = 718.3 kJ/kg × 1.5 kg/s
= 1077.45 kJ/s
= 1077.45 kW [as 1 W + 1 J/s] ... Ans.
Neglecting the changes in K.E. and P.E., the first law equation [Equation (1)] becomes
∆H = Q – Ws
– 761.07 = – 33.33 – Ws

∴ Ws = 727.74 kJ/kg

The power output of the turbine in kW is


·
Power output = Ws × m = 727.74 × 1.5 = 1091.61 kJ/s = 1091.61 kW ... Ans.
Chemical Engineering Thermodynamics - I 2.121 First Law of Thermodynamics

Example 2.49 :
A compressor is required to deliver air at 0.5 MPa and 200°C (473 K) to a stationary power
plant. The air intake is at 0.1 MPa and 20°C (293 K). The outlet velocity is 25 m/s. Calculate the
work per unit mass required for an adiabatic compressor.
Data : Cp for air = 1.0052 kJ/(kg·K) and is constant over the temperature range given.
Solution :
The first law of thermodynamics for a steady-state flow process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws … per unit mass of fluid … (1)

Assuming the potential energy change to be small compared with the other terms, we get
1 Out
∆H + 2 ∆u2 = Q – Ws

Since the compressor is adiabatic, Q = 0


1 Ws
∴ ∆H + 2 ∆u2 = – Ws … (2)

∆H = H2 – H1 Q=0
1 2 1 2 2
2 ∆u = 2 [u2 – u1] In
Fig. E 2.49
The inlet area to a compressor is often very large, so u1 ≈ 0 for a fixed mass flow rate.
1 2 1 2 1 2
2 ∆u = 2 [u2 – 0] = 2 u2
u2 = 25 m/s
1 2 1
∴ 2 ∆u = 2 (25) = 312.5 J/kg = 0.3125 kJ/kg
2

For an ideal gas, we have


∆H = Cp ∆T ... for unit mass of fluid
H2 – H1 = Cp (T2 – T1)
where Cp = 1.0052 kJ/(kg·K), T1 = 293 K and T2 = 473 K
∴ H2 – H1 = ∆H = 1.0052 (473 – 293) = 180.936 kJ/kg
Substituting the numerical values in Equation (2),
180.936 + 0.3125 = – Ws
∴ Ws = – 181.2485 kJ/kg ≈ – 181.25 kJ/kg
where the minus (–ve) sign indicates that the work is done on the system ⇒ work is required
by the compressor [it is a work consuming machine].
– Ws = 181.25 kJ/kg
∴ Work required by the adiabatic compressor –Ws = 181.25 kJ/kg ... Ans.
Chemical Engineering Thermodynamics - I 2.122 First Law of Thermodynamics

Example 2.50 :
A compressor to be purchased for the new mechanical engineering building must compress
air from 1 atm, 25°C (298 K) to 10 atm, 600°C (873 K). The outlet velocity from the compressor
must not exceed 10 m/s. Assume the compressor to be adiabatic.
Calculate : (i) the work required per unit mass for this compressor.
(ii) the power required to drive the compressor if 2 kg/s of air is to be compressed.
Data : Cp = 28.11 + 0.1967 × 10–2 T + 0.4802 × 10–5 T2 – 1.966 × 10–9 T3 kJ/(kmol·K)
Molecular weight of air = 29
Solution :
The first law of thermodynamics for a steady-state flow Out
process is given by
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid
Ws
… (1)
Units of each term ⇒ J/kg can be converted to kJ/kg Q=0
In this case, the potential energy change from inlet to outlet
is usually small, therefore
1 In
∆H + 2 ∆u2 = Q – Ws Fig. E 2.50
The inlet area of a compressor is often very large, so u1 = 0 for a fixed mass flow rate.
1 1 2 2 1 2
∴ 2 ∆u = 2 [u2 – u1] = 2 u2 , u2 = 10 m/s (given)
2

1
= 2 (10)2 = 50 J/kg = 0.05 kJ/kg
The compressor is adiabatic, so Q = 0. Equation (1) thus reduces to
1 2
∆H + 2 u2 = – Ws … (A)
∆H = H2 – H1
T2

For an ideal gas, ∆H = ⌠


⌡ Cp dT ... for unit mole of fluid
T1
where T1 = 298 K, T2 = 873 K
and Cp = 28.11 + 0.1967 × 10–2 T + 0.4802 × 10–5 T2 – 1.966 × 10–9 T3 kJ/(kmol·K)
873

∴ ∆H = ⌠
⌡ [28.11 + 0.1967 × 10–2 T + 0.4802 × 10–5 T2 – 1.966 × 10–9 T3] dT
298
0.1967 × 10–2 –– 2 –– 2
= 28.11 (873 – 298) + 2 [ 873 – 298 ]
0.4802 × 10–5 –– 3 –– 3 1.966 × 10–9 –– 4 –– 4
+ 3 [ 873 – 298 ] – 4 [ 873 – 298 ]
= 17566.48 kJ/kmol
Chemical Engineering Thermodynamics - I 2.123 First Law of Thermodynamics

Convert this ∆H in kJ/kg :


Molecular weight of air = 29 kg/kmol
17566 kJ/kmol
∴ ∆H = 29 kg/kmol = 605.74 kJ/kg

Substituting numerical values in Equation (A) yields


605.74 + 0.05 = – Ws
∴ Ws = – 605.79 kJ/kg
The minus sign indicates that the work is done on the system, i.e., work is required for the
compression of air.
The work required by the compressor (– Ws) is 605.79 kJ/kg ... Ans. (i)
·
Air flow rate = m = 2 kg/s
Power is work done per unit time
·
Power in kJ/s = (– Ws) in kJ/kg × m in kg/s
∴ Power required = [605.79 kJ/kg] [2 kg/s]
= 605.79 × 2 = 1211.58 kJ/s ≈ 1211.60 kW
Power required to drive the compressor = 1211.60 kW … Ans. (ii)
Example 2.51 :
1 kg/s of superheated steam at 2 bar and 673 K [H = 3276.6 kJ/kg] enters a turbine at a
velocity of 100 m/s. The turbine inlet is at an elevation of 10 m and the exit is at an elevation of
3 m. The steam leaves the turbine at a velocity of 150 m/s and is 98% dry at a pressure of 0.1 bar
[HV = 2584.70 kJ/kg and HL = 191.83 kJ/kg]. Calculate the power output of the turbine if the
energy loss from it is 40000 kJ/h.
Solution :
The first law of thermodynamics for a steady-state flow
process is
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... for unit mass of fluid … (1)
In
Units of each term ⇒ J/kg … can be converted to kJ/kg
∆H = H2 – H1
where H1 = H at the inlet = 3276.60 kJ/kg
Ws
H2 = H at the outlet
Quality of steam at the outlet = x = 0.98 (98% dry)
HL = 191.83 J/kg, HV = 2584.70 kJ/kg
H2 = xHV + (1 – x) HL
= 0.98 (2584.70) + (1 – 0.98) × 191.83 Out

= 2536.84 kJ/kg Fig. E 2.51


Chemical Engineering Thermodynamics - I 2.124 First Law of Thermodynamics

∴ ∆H = H2 – H1 = 2536.84 – 3276.60 = – 739.76 kJ/kg


u1 = Velocity at the inlet = 100 m/s
u2 = Velocity at the outlet = 150 m/s
1 1 2 2
2 ∆u = 2 [u2 – u1] ... is in J/kg
2

1
= 2 [(150)2 – (100)2] = 6250 J/kg = 6.25 kJ/kg

Heat loss from the turbine is 40000 kJ/h


40000
Heat loss = 40000 kJ/h = 3600 = 11.11 kJ/s

Since the heat is lost from the turbine (system) to the surroundings,
Q = – 11.11 kJ/s … by sign convention of Q
·
Mass flow rate of steam = m = 1 kg/s
Express Q in kJ/kg :
·
Q in kJ/kg = Q in kJ/s / m in kg/s
11.11 kJ/s
∴ Q = – 1 kg/s = – 11.11 kJ/kg

–ve sign indicates that the heat transfer is from the turbine (system) to the surroundings.
∆z = z2 – z1
z2 = Exit elevation of the turbine = 3 m
z1 = Inlet elevation = 10 m
∆z = 3 – 10 = – 7 m
g = 9.81 m/s2
∴ g ∆z = – 7 × 9.81 = – 68.67 J/kg = – 0.06867 ≈ – 0.0687 kJ/kg
Substituting numerical values in Equation (1),
– 739.76 + 6.25 + (– 0.0687) = – 11.11 – Ws
∴ Ws = 722.4687 ≈ 722.5 kJ/kg
Power is work done per unit time.
·
Power in kJ/s = Ws in kJ/kg × m kg/s
Power output of the turbine = 722.5 kJ/kg × 1 kg/s = 722.5 kJ/s = 722.5 kW ... Ans.
Example 2.52 :
Air is compressed from 1 bar and 27°C (300 K) to 30 bar in a two-stage compressor with
intercooling. Assume that the temperature of air leaving the intercooler to be 27°C (300 K) and
that the optimum interstage pressure is used. Calculate the work required per kg of air
compressed. Take n (polytropic exponent) = 1.3 and molecular weight of air = 29.
Chemical Engineering Thermodynamics - I 2.125 First Law of Thermodynamics

Solution :
Given : Optimum interstage pressure is used.
The optimum interstage pressure in a two-stage compressor is the geometric mean of the
initial and final pressures.
Let P' be the interstage (optimum) pressure.
P1 = Initial pressure = 1 bar
T1 = Initial temperature = 300 K
P2 = Final pressure = 30 bar
P' = (P1P2)1/2 = (1 × 30)1/2 = 5.477 bar
Polytropic exponent/index = 1.3
Molecular weight of air = 29 kg/kmol
For a multistage compressor and polytropic compression, the work of compression is given
by
Nn
Ws = n – 1 RT1 [1 – (r)(n–1)/n] ... kJ/kmol
where n = 1.3, N = number of stages = 2
R = 8.31451 kJ/(kmol·K), T1 = Initial temperature = 300 K
P' 5.477 P2 30
r = Compression ratio in each stage = P = 1 = 5.477 or r = P' = 5.477 = 5.477
1

Substituting numerical values yields,


2 × 1.3
Ws = (1.3 – 1) × 8.31451 × 300 [1 – (5.477)(1.3 – 1)/1.3]
= – 10389 kJ/kmol, i.e., –Ws = 10389 kJ/kmol
The minus (–ve) sign indicates that the work is done on the system, i.e., work is required for
compressing air.
Convert the work in kJ/kg.
Molecular weight of air = 29 kg/kmol
10389 kJ/kmol
Work required = – Ws = 29 kg/kmol = 358.24 kJ/kg
The work required per kg of air compressed is 358.24 kJ … Ans.
Example 2.53 :
A two-stage compressor is used to compress CO2 from its initial state of 0.5 bar and
27°C (300 K) to a final pressure of 1.5 bar with intercooling to 300 K. A compression efficiency
of 85% may be assumed in each stage. Calculate the work required per kg of CO2 for driving the
compressor and the discharge temperature. Take γ = 1.3, Mol. wt. of CO2 = 44.
Solution :
Given : After first stage, the temperature of the gas is brought down to the initial value
(300 K).
Chemical Engineering Thermodynamics - I 2.126 First Law of Thermodynamics

∴ For a multistage compressor with n stages and with the same pressure ratio in each
stage, the work required is given by

Ws =
γ–1
[
RT1 1 – (r)
(γ–1)/γ
]
where r = Compression ratio in each stage

= P 
P2 1/N
 1
N = no. of stages = 2, P1 = 0.5 bar (initial pressure), P2 = 1.5 bar (final pressure)

r = 0.5 = 1.732
1.5 1/2

 
or P' = Interstage pressure (optimum)
The optimum interstage pressure in a two-stage compressor is the geometric mean of the
initial and final pressures.
P' = (P1P2) = (0.5 × 1.5) = 0.866 bar
P' 0.866 P2 1.5
∴ r = P = 0.5 = 1.732 = P' = 0.866 = 1.732
1

γ = 1.3, R = 8.31451 kJ/(kmol·K), T1 = initial temperature = 300 K


2 × 1.3
Ws = (1.3 – 1) × 8.31451 × 300 [1 – (1.732)(1.3 – 1)/1.3]
= – 2921.4 kJ/kmol
where the minus (–ve) sign signifies that the work is done on the system.
∴ Work required,
– Ws = 2921.42 kJ/kmol
Molecular weight of CO2 = 44 kg/kmol
Work required per kg of CO2 is
2921.42 kJ/kmol
Work required (theoretical) = 44 kg/kmol = 66.39 kJ/kg
Compression efficiency = 85%, i.e., η = 0.85
66.39
Actual work required = 0.85 = 78.1 kJ/kg ... Ans.
Temperature and pressure in an adiabatic compression are related by
(γ–1) / γ
P2
T2 = T1 P 
 1
(γ–1) / γ
In our case,
P2
T2 = T' P' , T2 = Discharge temperature
 
where T' = 300 K, P2 = 1.5 bar, P' = 0.866 bar

T2 = 300 0.866
1.5 (1.3 – 1)/1.3
∴ = 340.55 K ... Ans.
 
Chemical Engineering Thermodynamics - I 2.127 First Law of Thermodynamics

Example 2.54 :
A two-stage compressor with intercooling is used to compress an ideal gas. The ideal gas at
T1 and P1 enters the first stage of the compressor and leaves at P. The gas from the first stage is
then cooled isobarically to its initial temperature T1 in the intercooler. It is then compressed to
P2 in the second stage of the compressor.
(i) Derive a relation to calculate the work to be done per mole of the gas.
(ii) Determine the pressure P (to which the gas is to be compressed in the first stage) for
which the total work done on the compressor is minimum.
Solution :
Part (i) :
P, T T1, P
Intercooler

Stage - 1 Stage - 2

P1, T1 P2, T2

Fig. E 2.54
The work required for compressing an ideal gas adiabatically from P1 to P is given by
 P (γ–1) / γ
P1V1 1 – P 
γ
(Ws)1 = 
γ–1   1 
Similarly, for stage-2 from P to P2,
γ  P2(γ–1) / γ
(Ws)2 = PV 1 – P 
γ–1    
The work required in a multi-stage compressor is the sum of the work required for the
individual stages.
For a two-stage compressor in which an ideal gas is compressed from P1 to P in the first stage
and cooled to T1 in the intercooler and then compressed to P2, the total work required is given by
Ws = (Ws)1 + (Ws)2
γ–1  P (γ–1) / γ  P2(γ–1) / γ
P1V1 1 – P 
γ
Ws =  + PV 1 – P 
γ   1  γ–1    
The amount of gas compressed is the same in both the stages.
For an ideal gas,
PV = nRT
Chemical Engineering Thermodynamics - I 2.128 First Law of Thermodynamics

At the inlet to the first stage, it becomes,


P1V1 = nRT1
At the inlet to the second stage, it becomes
PV = nRT1
∴ P1V1 = PV = nRT1
i.e., P1V1 = PV
 P (γ–1) / γ  P2(γ–1) / γ
P1V1 1 – P 
γ γ
∴ Ws =  + P1V1 1 – P 
γ–1   1  γ–1    
 P (γ–1) / γ P2(γ–1) / γ
P1V1 2 – P 
γ
= – P 
γ–1   1   
For 1 mole of an ideal gas, P1V1 = RT1
 P (γ–1) / γ P2(γ–1) / γ
RT1 2 – P 
γ
∴ Ws = – P  ... in kJ/kmol or J/mol ... Ans. (i)
γ–1   1   
If R is in kJ/(kmol·K), then Ws is in kJ/kmol.
The above equation is the desired relation/expression to calculate the work done per mole of
an ideal gas for a two-stage compressor.
Part (ii) : In the above expression, P1 and P2 are fixed. The only variable pressure for
minimum total work is P.
dWs
The condition for minimum work required : dP = 0
To determine the pressure P, for which the total work is minimum, differential work done
dWs
with respect to P and equate the derivative  dP  obtained to zero.
 
γ RT1   P (γ–1) / γ P2(γ–1) / γ
Ws = 2 – – P 
γ – 1  P1   
γ–1
Let = x
γ
γ
Let RT1 = C = constant
γ–1
 P x P2x
∴ Ws = C 2 – P  – P 
  1  
d(Px) x–1
We have, dP = xP
d(1/Px) d(P–x) –x–1 = –
x
dP = dx = – xP x
P +1
dWs  xPx–1 xP2  x
= C 0 – 
dP  x + Px+1
P 1 
Chemical Engineering Thermodynamics - I 2.129 First Law of Thermodynamics

dWs
Setting dP = 0, we get

 xPx–1 xPx2 
0 = C 0 – x + x+1
 P 
 P1 
x
xPx–1 xP2
x = x+1
P1 P
x x
Px–1 · Px+1 = P1 P2
P2x = (P1P2)x
∴ P = P1P2 ... Ans. (ii)
That is, the optimum interstage pressure (i.e., the interstage pressure for which the total work
required is minimum) in a two-stage compressor is the geometric mean of the initial and final
pressures.
Extra : P = (P1P2)1/2
Dividing the above relation by P1,
P (P1P2)1/2 P21/2
= = P  ... Pressure ratio in the first stage
P1 P1  1
Similarly, dividing by P2,
P (P1P2)1/2 P11./2
= = P 
P2 P2  2
P2 P21/2
∴ = P  ... Pressure ratio in the second stage
P  1
∴ For total minimum work in a two-stage compressor, the pressure ratios in both stages
are equal and are equal to the square root of the overall pressure ratio.
Substituting the value of P in the work expression, we get
 P (γ–1) / γ P2(γ–1) / γ
RT1 2 – P 
γ
Ws = – P  … work expression
γ–1   1   
γ  (P1P2)1/2(γ–1) / γ  P2 (γ–1) / γ
= RT1 2 –  P  – (P P )1/2 
γ–1   1   1 2  
 P2 (γ–1) / 2γ P2(γ–1) / 2γ
RT1 2 – P 
γ
= – P  
γ–1   1  1 
2γγ  P2(γγ–1) / 2γ
Ws = RT1 1 – P  
γ–1   1 
This is the desired expression for the minimum total work in a two-stage compressor.
P2 1/2
Compression ratio in both the stages = P 
 1
Chemical Engineering Thermodynamics - I 2.130 First Law of Thermodynamics

∴ For a multistage compressor with N stages, Ws is given by


 P2 (γ–1) / Nγ
RT1 1 – P 

Ws = 
γ–1   1 
1/N
Compression ratio in all the stages = P 
P 2
 1
For a three-stage compressor with inter-stage cooling :
P2 P3 P4 P41/3
= = = P 
P1 P2 P3  1
P2 = P1 P 
P4 1/3

 1
P3 = P2 P  = P1 P  P 
P4 1/3 P4 1/3 P4 1/3
 1  1  1
P3 = P1 P 
P4 2/3

 1
Example 2.55 :
A two-stage single acting air compressor draws 6 kg/min air at 1 bar and 20°C (293 K). The
delivery pressure is 15 bar. Assume perfect intercooling. Take polytropic exponent / compression
index, n = 1.25. Clearance factors in low pressure and high pressure cylinders are 0.07 and 0.04
respectively. Speed of the compressor is 420 r.p.m.
Determine : (a) the actual power required when the compressor efficiency is 85%, (b) the
isothermal efficiency, (c) the volumetric efficiency in each stage and (d) the heat transferred in
the intercooler.
Take R = 8.31451 kJ/(kmol·K), Cp = 1.005 kJ/(kg·K) and Cv = 0.718 kJ/(kg·K),
R = 0.287 kJ/(kg·K), Mol. wt. of air = 29
Solution :
P1 = 1 bar ... pressure at the inlet
P6 = 15 bar ... delivery pressure
T1 = 293 K ... temperature at the inlet
n = 1.25
N = number of stages = 2
.
m = 6 kg/min = 0.1 kg/s
For a two-stage compressor (with inter-
cooling) the work of compression is given by
N·n (n–1)/n
Ws = (n – 1) RT1 [1 – (r) ] ... kJ/kmol Fig. E 2.55

where r = compression ratio in each stage =  P at inlet  = P  =  1  = 3.873


P at delivery 1/N P6 1/N 15 1/2
   1  
Chemical Engineering Thermodynamics - I 2.131 First Law of Thermodynamics

T1 = 293 K, R = 8.31451 kJ/(kmol·K), N = 2 and n = 1.25


2 × 1.25 (1.25 – 1) / 1.25
∴ Ws = (1.25 – 1) × 8.31451 × 293 [1 – (3.873) ]
= – 7576.93 kJ/kmol
Work required
∴ by the compressor = (– Ws) = 7576.93 kJ/kmol
Molecular weight of air = 29 kg/kmol
·
m‚ kg/s
Molar flow rate of air = Mol. Wt. of N ‚ kg/kmol
2

0.1
= 29 = 3.45 × 10–3 kmol/s

Power required in kJ/s = (–Ws) in kJ/kmol × Molar flow rate of air in kmol/s
Power required (theoretical) = 7576.93 × 3.45 × 10–3 = 26.14 kJ/s = 26.14 kW
Compressor efficiency = 85% or η = 0.85
Theoretical power
∴ Actual power required =
η
26.14
= 0.85 = 30.753 ≈ 30.75 kW ... Ans. (a)

Isothermal work required = RT1 ln P 


P6
 1

= 8.31451 × 293 ln  1  = 6597.2 kJ/kmol


15
 
= 6597.2 × 3.45 × 10–3 = 22.76 kJ/s = 22.76 kW
Isothermal power
Isothermal efficiency = Polytropic power

22.76
= 26.14 = 0.871 or 87.10% ... Ans. (b)
Volumetric efficiency (LP cylinder - first stage)
V1 – V4 V1 – V4 P21/n
= V –V = = 1 + C – C P 
1 3 V3  1
V3 Vc
where C = Clearance = V – V = V = 0.07
1 3 s
C = Clearance = Clearance factor = Clearance ratio = 0.07
For perfect intercooling, P2 = 1 × 15 = 3.873 bar
P1P6 =
Volumetric efficiency = 1 + 0.07 – 0.07  1 
3.873 1/1.25
 
= 0.863 or 86.30% ... Ans. (c)
Chemical Engineering Thermodynamics - I 2.132 First Law of Thermodynamics

P61/n
Volumetric efficiency (high pressure cylinder) = 1 + C – C P 
 5
P5 = P2 = Intermediate pressure = 3.873 bar
C = 0.04

Volumetric efficiency = 1 + 0.04 – 0.04 3.873


15 1/1.25
 
= 0.9218 or 92.18% ... Ans. (c)
P2 P61/N = 151/2 = 3.873
We have : = P  1
P1  1
P2(n–1)/n
T2 = T1 P 
 1
(1.25 – 1)/1.25
= 293 (3.873)
= 384.13 ≈ 384 K
Heat transfer in the intercooler where air is cooled isobarically from 384 K to 293 K
(temperature at the inlet to second stage) is given by
.
Q = m Cp (T1 – T2)
.
where m = 0.1 kg/s ... Mass flow rate of air
Cp = 1.005 kJ/(kg·K)
T2 = 384 K = T at inlet to intercooler
and T1 = 293 K = T at outlet to intercooler
Q = 0.1 × 1.005 (293 – 384)
= – 9.145 kJ/s = – 9.145 kW
–ve sign indicates that heat is removed in the intercooler.
Heat to be removed in the intercooler = 9.145 kJ/s = 9.145 kW ... Ans. (d)
Example 2.56 :
A small water pump located 15 m down in a well takes water in at 283 K, 90 kPa at a rate of
1.5 kg/s. The exit line of the pump is a pipe of 40 mm diameter that goes upto a receiving tank
maintaining a pressure of 400 kPag. Assume that the process to be adiabatic, the same inlet and
exit velocities and the water stays at 283 K. Find the pump work required.
Solution :
The first law of thermodynamics for a steady flow process gives
1
∆H + 2 ∆u2 + g ∆z = Q – Ws ... per kg of water
The process is adiabatic (given), so Q = 0
For the same inlet and exit velocities : ∆u2 = 0
Chemical Engineering Thermodynamics - I 2.133 First Law of Thermodynamics

The first law equation reduces to


∆H + g ∆z = – Ws
We have : H = U + PV
∆H = H2 – H1 = (P2 – P1) V
... (since V is constant and U is constant)
V = Specific volume of water
1 1
= = 1000 = 0.001 m3/kg
ρ of water
P1 = 90 kPa, P2 = 400 kPag = 400 + 101.3 = 501.3 kPa
[P2 (absolute) = P2 gauge pressure + atmospheric pressure]
Let z2 = 15 m, then z1 = 0 m
∴ z2 – z1 = 15 – 0 = 15 m
g = 9.81 m2/s2
g ∆z = 9.81 m/s2 × 15 m = 147.15 m2/s2
m2·kg kg·m m N·m
= 147.15 s2·kg = 147.15 s2 · kg ⇒ kg
147.15
= 147.15 J/kg = 1000 kJ/kg

2 P = 400 kPag
T = 283 K

15 m

1
283 K
90 kPa
Fig. E 2.56
We have : ∆H + g ∆z = – Ws
(P2 – P1) V + g (z2 – z1) = – Ws
147.15
(501.3 – 90) × 0.001 + 1000 = – Ws
∴ –Ws = 0.55845 kJ/kg
∴ Ws = – 0.55845 kJ/kg
The minus sign signifies that the work is to be done on the system (pump).
.
Mass flow of water = m = 1.5 kg/s
∴ Work required by the pump (– Ws) = 0.55845 kJ/kg
Chemical Engineering Thermodynamics - I 2.134 First Law of Thermodynamics

Power is work done per unit time


·
Power required in kJ/s = (–Ws) in kJ/s × m in kg/s
Power required = 0.55845 × 1.5 = 0.83765 ≈ 0.84 kJ/s = 0.84 kW
∴ The pump requires a power input of 0.84 kW, i.e., 840 W ... Ans.
Example 2.57 :
A pressure cylinder of volume V contains air at pressure Po and temperature To. It is to be
filled from a compressed air line maintained at constant pressure P1 and temperature T1. Show
that the temperature of the air in the cylinder after it has been charged to the pressure of the line
is given by
γT1
T = Po  T1
1 + P γ T – 1

1 o 
Solution :
Volume of the cylinder = V
Initial air pressure in the cylinder = Po
Initial temperature of the air in the cylinder = To
Pressure of the compressed air line = P1
Temperature of the air in the compressed air line = T1
Temperature of the air in the cylinder after it is charged with air from the compressed air
line = T.
Pressure of the air in the cylinder after charging = pressure of the compressed air line = P1.
In the absence of any information regarding heat transfer during the charging process,
assume that the gas is suddenly charged so that there is no enough time for the cylinder to
exchange heat with the surroundings. That is Q = 0.
The first law of thermodynamics for the charging process gives
Final energy = Initial energy + due to air flow +  to heat transfer 
Energy added Energy added due

With Q = 0, it becomes
Final energy – Initial energy = Energy added due to air flow.
nf Uf – no Uo = (nf – no) H … (A)
where H = enthalpy of the air in the compressed air line.
Uo = initial internal energy of the air in the cylinder
Uf = final internal energy of the air in the cylinder
nf = final moles of the air in the cylinder = PfV/RTf
Pf = final pressure of the air in the cylinder = P1 = pressure of the
compressed air line
Tf = final temperature of the air in the cylinder = T
Uf = Cv Tf = CvT, H = CpT1 and Uo = Cv To … for an ideal gas
no = initial moles of the air in the cylinder
Chemical Engineering Thermodynamics - I 2.135 First Law of Thermodynamics

For an ideal gas : PV = nRT


PV
n = RT , V = volume of the cylinder

∴ no = PoV/RTo and nf = PfV/RTf = P1V/RT


(since Tf = T and Pf = P1)
Substituting the values of the terms involved in Equation (A), we get
P1V PoV P1V PoV
RT CvT – RTo CvTo =  RT – RTo Cp T1
On simplification, it becomes

Cv P1 – Cv Po =  T – T  Cp T1
P1 Po
 o

Dividing both sides by Cv gives


P1 Po  Cp
P1 – Po =  T – T  C T1
 o v
Cp
Since Cv = γ, the above equation becomes
P1 Po
P1 – Po = T γ T1 – T γ T1
o

Rearranging gives
Po P1γ T1
P1 – P0 + T γ T1 = T
o
P1 γ T1
∴ T = T1
P1 – Po + Po · γ T
o
Dividing the numerator and denominator of the RHS term by P1 gives
γ T1
T = Po Po T1
1–P +P γ T
1 1 o
Taking +Po/P1 common from the second and third terms of the numerator on the RHS, we get
γT1
T =
1 + P γ T – 1
Po T1
1  o 
Example 2.58 :
A stream of gases at 7.5 bar and 1023 K (750°C) is passed through a turbine of a jet engine
at a velocity of 140 m/s. The stream leaves the turbine at 2 bar, 823 K and 280 m/s. The turbine
may be assumed to be adiabatic. The enthalpy of gases at the inlet and exit of the turbine is
950 kJ/kg and 650 kJ/kg, respectively.
Find the capacity of the turbine if the gas flow rate is 5 kg/s.
Chemical Engineering Thermodynamics - I 2.136 First Law of Thermodynamics

Solution :

2 2

Turbine WS

Adiabatic, Q = 0
1 1

Fig. E 2.58
The turbine is adiabatic, so that Q = 0.
For a steady-state flow process, the first law of thermodynamics is
∆u2
∆H + 2 + g ∆z = Q – Ws … J/kg
With g ∆z = 0 and Q = 0, the first law equation becomes
2 2
(u2 – u1)
(H2 – H1) + 2 = –Ws … (1)
Ws = work produced by the turbine = work output of the turbine
H1 = 950 kJ/kg = 950 × 103 J/kg, H2 = 650 × 103 J/kg
u1 = 140 m/s and u2 = 280 m/s
Substituting these values in Equation (1), we get
(280)2 (140)2
(650 × 103 – 950 × 103) + 2 – 2 = – Ws
– Ws = – 270600 J/kg
∴ Ws = 270600 J/kg = 270.6 kJ/kg
·
Mass flow rate of gases = m = 5 kg/s
.
Power output in kJ/s = Ws in kJ/kg × m in kg/s
Power capacity
.
(power output) = m · Ws
of the turbine
= 5 kg/s × 270.6 kJ/kg
= 1353 kJ/s
= 1353 kW … Ans.
Example 2.59 :
In a gas turbine, the gas enters at the rate of 5 kg/s, with a velocity of 50 m/s and enthalpy
of 900 kJ/kg, at 100 kPa and 300 K (27°C). It leaves the turbine with a velocity of 150 m/s and
enthalpy of 400 kJ/kg. Assuming the heat loss from the gas to the surroundings to be 25 kJ/kg,
determine the power output of the turbine and the diameter of the inlet pipe.
Assume ideal behaviour and take R = 0.285 kJ/(kg.K).
Chemical Engineering Thermodynamics - I 2.137 First Law of Thermodynamics

Solution :

2 2

Turbine WS

1 1 Q

Fig. E 2.59
The first law of thermodynamics for a steady-state flow process is
∆u2
∆H + 2 + g ∆z = Q – Ws … in J/kg … per unit mass of fluid

Neglecting g ∆z, it becomes


2 2
u2 u1
(H2 – H1) +  2 – 2  = Q – Ws … (1)
 
H2 = 400 kJ/kg = 400 × 103 J/kg
H1 = 900 kJ/kg = 900 × 103 J/kg
u2 = 150 m/s, u1 = 50 m/s
Q = Heat loss from the gas
Q = heat transferred to the gas from the surroundings
= – heat loss from the gas to the surroundings
= – 25 kJ/kg = – 25 × 103 J/kg … by sign convention
Ws = work output of the turbine/work produced in J/kg
Substituting the values of the various terms in Equation (1), we get
(150)2 (50)2
(400 – 900) × 103 + 2 – 2 = – 25 × 103 – Ws
∴ – 465000 = – Ws
∴ Ws = 465000 J/kg … work produced/delivered by the turbine
·
Mass flow rate = m = 5 kg/s
·
Power output in J/s = m in kg/s × Ws in J/kg
Power output ·
of the turbine = m Ws
= 5 × 465000 = 2325000 J/s
= 2325 kJ/s = 2325 kW … Ans.
Chemical Engineering Thermodynamics - I 2.138 First Law of Thermodynamics

Let us find the diameter of the inlet pipe.


For an ideal gas,
PV = n RT
or PV = m RT, where m = mass of the gas
· ·
PV = m RT
·
where, V = volumetric flow rate, m3/s
R = 0.285 kJ/(kg.K) = 0.285 × 103 J/(kg.K)
P = 100 kPa = 100 × 103 Pa and T = 300 K
· ·
mRT
V = P
5 × 0.285 × 103 × 300
=
100 × 103
= 4.275 m3/s
u1 = velocity at the inlet = 50 m/s
π
Cross-sectional area = A = 4 D2
· π
V = u1A = u1 4 D2

 4V· 1/2
4 × 4.2751/2
∴ D =   =  
πu1  π × 50 
= 0.3299 ≈ 0.33 m
Diameter of the inlet pipe = 0.33 m or 33 cm … Ans.
Example 2.60 :
In a gas turbine, the gas enters at the rate of 15 kg/s with a velocity of 50 m/s and enthalpy
of 1260 kJ/kg and leaves the turbine with a velocity of 110 m/s and enthalpy of 400 kJ/kg.
The power developed by the turbine is 12000 kW.
Calculate the rate at which heat is rejected from the turbine and the area of the inlet pipe.
Data : Specific volume of the gas at the inlet is 0.45 m3/kg.
Solution :

2 2
Q

WS

1 1

Fig. E 2.60
Chemical Engineering Thermodynamics - I 2.139 First Law of Thermodynamics

Power developed by the turbine = 12000 kW = 12000 × 103 W (J/s)


·
Mass flow rate of gas = m = 15 kg/s
Power developed 12000 × 103
Ws = work produced by the turbine = Mass flow rate = 15
= 8 × 105 J/kg
H1 = enthalpy of gas at the inlet = 1260 kJ/kg = 1260 × 103 J/kg
H2 = enthalpy of gas at the outlet = 400 kJ/kg = 400 × 103 J/kg
u1 = velocity at the inlet = 50 m/s
u2 = velocity at the outlet = 110 m/s
For a steady-state flow process, the first law gives
∆u2
∆H + 2 + g ∆z = Q – Ws … energy per kg ⇒ J/kg
Neglecting g ∆z, it becomes
2 2
u2 – u1
(H2 – H1) +  2  = Q – Ws
 
Q = Heat supplied/added
Ws = Work output (work produced) from the turbine
Substituting the values, we get
(110)2 (50)2
(400 × 103 – 1260 × 103) + 2 – 2 = Q – 8 × 105

– 855200 + 8 × 105 = Q
Q = – 55200 J/kg = – 55.20 kJ/kg
where the –ve sign indicates that the heat is rejected from the turbine, i.e., there is a heat loss
from the turbine to the surroundings.
Heat rejected/lost (–Q) = 55.20 kJ/kg
Heat rejected from the turbine = 55.20 kJ/kg … Ans.
·
(– Q) in kJ/s = (– Q) in kJ/kg × m in kg/s
Heat rejected/Heat loss (– Q)
from the turbine = 55.20 kJ/kg × 15 kg/s = 282 kJ/s = 828 kW … Ans.
Calculation of the inlet area :
mass flow rate = density × velocity × area
.
m = ρ · us · A
.
∴ A = m /ρu1
.
A = m × specific volume/u1 [since density = 1/specific volume]
Inlet area = A = 15 × 0.45/50 = 0.135 m2 … Ans.
Chemical Engineering Thermodynamics - I 2.140 First Law of Thermodynamics

Example 2.61 :
A heat exchanger is to be designed to raise the temperature of ammonia, flowing through the
exchanger at a rate of 10 kmol/h, from 27°C (300 K) to 127°C (400 K). Find the amount of heat
to be transferred to the ammonia in the exchanger.
Data : Cp for ammonia is given by
Cp = 29.747 + 25.108 × 10–3 T – 1.546 × 105 T–2 , J/(mol.K)
Solution :
For a heat exchanger :
The KE and PE terms are considered to be very small, so that ∆u2/2 = 0 and g∆z = 0
(compared to enthalpy). There is no external work done, so that Ws = 0. There is no heat loss
(insulated) and there is only heat transfer between ammonia and a heating medium.
The first law equation for a steady-state flow process is
∆u2
∆H + g∆z + 2 = Q – Ws
With the above mentioned conditions, it reduces to
Q = ∆H
The inlet and exit ammonia pressures are almost identical, so it is a constant pressure heating
process.
For this constant pressure process, ∆H is given by
T2

∆H = ⌠
⌡ Cp dT, Given : T1 = 300 K and T2 = 400 K
T1
T2 T2

∴ Q = ⌠
⌡ Cp dT = ⌠
⌡ (29.747 + 25.108 × 10 T – 1.546 × 10 T ) dT
–3 5 –2

T1 T1

× 10–3 (T2 – T1 ) + 1.546 × 105 T – T 


25.108 2 2 1 1
= 29.747 (T2 – T1) + 2  2 1

[(400)2 – (300)2] + 1.546 × 105 400 – 300


25.108 1 1
= 29.747 (400 – 300) + 2
 
= 37246500 J/h = 37.246 MJ/h
∴ Amount of heat to be transferred to ammonia = 37.246 MJ/h … Ans.
Example 2.62 :
To relate P and V of any reversible process, an empirical equation PVδ = c, where c is a
constant, is used.
Show that for an ideal gas, this equation leads to
RT1  P2 δ 
δ–1
W = 1 –   
δ – 1  P1 
If the process is isothermal, δ = 1, show that the above equation reduces in this case to
P1
W = RT ln P
2
Chemical Engineering Thermodynamics - I 2.141 First Law of Thermodynamics

Solution :
For any reversible process, the work done is given by
W = ⌠
⌡ P dV
Let P1, V1 and T1 be the pressure, volume and temperature, respectively in the initial state of
the gas and P2, V2 and T2 be the corresponding values in the final state.
For a reversible process : PVδ = c is valid (given)
c
∴ P = δ
V
With this, the work expression becomes
V2
1
W = c⌠
⌡ Vδ dV
V 1
V2 V2
 V–δ+1 
W = c⌠
⌡ V–δ dV = c  
V1
– δ + 1V1
c δ δ –c δ δ
= [V1– – V1– ] = [V1– – V1– ]
1–δ 2 1 1–δ 1 2

1 δ δ
W = [cV1– – cV1– ] … (1)
δ–1 1 2

We have PVδ = c ∴ P1Vδ1 = c and P2Vδ2 = c … (2)


Replacing c from Equation (1) by its values in Equation (2) gives
P1Vδ1 V1–
1
δ
– P2Vδ2 V1–
2
δ
P1V1 – P2V2
W = = … (3)
δ–1 δ–1
For an ideal gas : PV = RT … for 1 mole
∴ P1V1 = RT1 and P2V2 = RT2 … (4)
Combining Equations (3) and (4), we get
RT1 – RT2
W =
δ–1
RT1  T2
= 1 – 
δ – 1  T 1
… (5)

We have, PVδ = c … (6)


For an ideal gas : PV = RT
RT
∴ V = P … (7)
Chemical Engineering Thermodynamics - I 2.142 First Law of Thermodynamics

Substituting V from Equation (7) in Equation (6), we get


RT δ
P P  = c
 
 Tδ 
R  δ–1 = c
P 

i.e., = constant … [since c and R are constants]
Pδ–1
T
or = constant
P(δ–1)/δ
δ–1
δ
or T = constant × P
∴ T1 = constant × P(δ
1
–1)/δ

T2 = constant × P(δ
2
–1)/δ

δ–1
T2 P2 δ
∴ T1 = P1 … (8)

Combining Equations (5) and (8), we get


 P2 δ 
δ–1
W =
RT1 
1 –    … (1) … Ans.
δ–1  P1 
Part II : For an isothermal process :
P1
W = RT ln P … (2)
2

We have to obtain Equation (2) from Equation (1).


Given : For an isothermal process, δ = 1.
δ–1
RT1  P2 δ 
δ–1 δ
– RT1 [1 – (P2/P1) ]
We have : W = 1–   = … (3)
δ – 1  P1  [1 – δ]
0
If we put δ = 1 in the above equation, the RHS will be of 0 form (indeterminate form).

[If f(x) and g(x) are two functions of x such that f(a) = g(a) = 0
lim f(x) 0
x → a g(x) = 0 , then use L'Hospital rule.
lim f(x) lim f'(x) f'(a) 0
= x → a g'(x) = g'(a) . If it is of 0 form, then again
x → a g(x)
differentiate and put x = a and so on].
Chemical Engineering Thermodynamics - I 2.143 First Law of Thermodynamics

In our case, x = δ and a = 1.


Let us apply L'Hospital rule to Equation (3).
  δ–1  
lim  d  P2 δ 
(1 – δ)
d
W = – RT1 δ → 1 1– 
dδ  P1 
/ dδ 
… (4)

Evaluate the derivatives separately :


d
(1 – δ) = 0 – 1 = – 1


d  P2 δ 
 δ–1 δ–1
P2 d δ –1
= 0 – P  ln P 
P2 δ
1 – P   
dδ   1   1  1 dδ  δ 
δ–1
= 0 – P  · ln P  0 + 2
P2 1 P2 1
 1  1  δ 
δ–1
= – P  ln P  · 2
P2 δ P2 1
 1  1 δ
For obtaining the above derivative, we have used :
d f(x) f(x)
dx (a ) = a · ln (a) · f'(x)
Substituting the values of the derivative terms, Equation (4) becomes
 δ–1 
δ
lim – (P2/P1) · ln (P2/P1) · 1/δ  2
W = –RT1 δ → 1  
 –1
 δ–1 
δ
lim (P2/P1) · ln (P2/P1) · 1/δ2
= – RT1 δ → 1  
 1
Putting δ = 1, the above equation becomes
1–1
P2 1 P2 1
W = –RT1 P  · ln P  · (1)2
 1  1
= –RT1 × 1 × ln P  · 1
P2
 1
= –RT1 [ln P2 – ln P1]
= RT1 [ln P1 – ln P2]
For an isothermal process : T = constant ∴ T1 = T
∴ W = RT [ln P1 – ln P2]
P1
W = RT ln P  … Ans.
 2
Chemical Engineering Thermodynamics - I 2.144 First Law of Thermodynamics

Example 2.63 :
Derive an expression for the work of reversible, isothermal compression of 1 mol of a gas
from an initial volume V1 to a final volume V2.
The equation of state is
P (V – b) = RT, where b is a positive constant
Solution :
For a reversible, isothermal process, the work done is given by
W = ⌠
⌡ P dV … (1)
The equation of state is
P (V – b) = RT
RT
P = V–b … (2)
Combining Equations (1) and (2) gives
RT dV
W = ⌠⌡ V – b dV = RT ⌠
⌡V–b
since T is constant for an isothermal process.
Integrating at constant T from the initial volume V1 to the final volume V2 gives
V2
dV
W = RT ⌠
⌡ V–b
V1

V2 – b
W = RT ln V – b … Ans.
 1 
Example 2.64 :
1 kmol of oxygen undergoes a reversible non-flow isothermal compression and the volume
decreases from 0.20 m3/kg to 0.08 m3/kg. The initial temperature of the gas is 333 K (60°C).
If the gas obeys the van der Waals equation, calculate the work done during the process and the
final pressure.
Data : For O2 gas, a = 1.3925 × 105 (N.m4)/kmol2
and b = 0.0314 m3/kmol
The van der Waals equation of state is
P + a  (V – b) = RT
 V2
Solution :
Initial volume = 0.20 m3/kg, Final volume = 0.08 m3/kg
Mol. Wt. of O2 = 32 kg/kmol
Initial molar volume = V1 = 0.20 × 32 = 6.4 m3/kmol
Final molar volume = V2 = 0.08 × 32 = 2.56 m3/kmol
Initial temperature = T1 = 333 K = T
Amount of O2 gas = 1 kmol
Chemical Engineering Thermodynamics - I 2.145 First Law of Thermodynamics

Work done per 1 kmol of O2 is given by


2
⌠ P dV
W = ⌡ … (1)
1

The van der Waals equation is


P + a  (V – b) = RT
 V 2
RT a
∴ P = V – b – V2 … (2)
Replacing P from Equation (1) by Equation (2) gives
V2 V2 V2

W = ⌠  RT a  RT  a
⌡ V – b – V2 dV = ⌠
⌡ V – b dV – ⌠
⌡ V2 dV
V1 V1 V1

Since the process is isothermal, T is constant.


V2

+ a V
V2 1
W = RT [ln (V – b)]V1
 V1
V2 – b
= RT ln V – b + a V – V 
1 1
 1   2 1

R = 8.31451 N.m/(mol.K) = 8314.51 N.m/(kmol.K)


T = 333 K, a = 1.3925 × 105 (N.m4)/kmol2, b = 0.0314 m3/kmol
W = 8314.51 × 333 ln  6.4 – 0.0314  + 1.3925 × 105 2.56 – 6.4
2.56 – 0.0314 1 1
   
= – 2557516 + 32636.7 = – 2524879.3
= – 2524879 (N.m)/kmol … Ans.
The minus sign signifies that the work is done on the gas.
Final pressure = P2
RT a
P2 = V – b – 2
2 V2
8314.51 × 333 1.3925 × 105
= 2.56 – 0.0314 – (2.56)2
= 1073718 N/m 2 (105 N/m2 = 1 bar)
= 10.7318 bar ≈ 10.73 bar … Ans.
Example 2.65 :
The enthalpy of a fluid at the inlet to a certain nozzle is 2800 kJ/kg. The velocity at the inlet
is 50 m/s. At the discharge end (exit), the enthalpy is 2600 kJ/kg.
(a) Find the velocity at the exit of the nozzle.
(b) Find the mass flow of the fluid if the inlet area is 900 cm2 and the specific volume at the
inlet is 0.187 m3/kg.
(c) If the specific volume at the exit of the nozzle is 0.498 m3/kg, find the exit area of the
nozzle.
Assume that the nozzle is horizontal and that the heat loss from the nozzle is negligible.
Chemical Engineering Thermodynamics - I 2.146 First Law of Thermodynamics

Solution :
At the inlet : H1 = 2800 kJ/kg = 2800 × 103 J/kg,
u1 = 50 m/s, A1 = 900 cm2 = 900 × 10–4 m2
·
v1 = specific volume = 0.187 m3/kg, m = ?
At the exit : H2 = 2600 kJ/kg = 2600 × 103 J/kg, u2 = ?, A2 = ?
v2 = 0.498 m3/kg
The first law of thermodynamics for a steady flow process gives
1
∆H + 2 ∆u2 + g ∆z = Q – Ws … J/kg … (1)
(i) There is no shaft work delivered in a nozzle. Therefore, Ws = 0.
(ii) The nozzle is horizontal. Therefore, ∆z = 0.
(iii) The heat loss from the nozzle is negligible. Hence, Q = 0.
With these conditions, Equation (1) reduces to
1
∆H + 2 ∆u2 + 0 = 0 – 0
1
∆H + 2 ∆u2 = 0
1 2 1 2 2 1 2 1 2
2 ∆u = 2 (u2 – u1) = 2 u2 – 2 u1 = – ∆H = – (H2 – H1)
Substituting the values,
1 2 1
2 u2 – 2 (50) = – (2600 – 2800) × 10
2 3

2
u2 = 402500 m2/s2, u2 = (402500)1/2 = 634.43 m/s
The velocity at the exit of nozzle is 634.43 m/s … Ans. (a)
·
Calculation of the mass flow rate (m ) :
·
m = ρuA = ρ1u1A1 = ρ2u2A2 … continuity equation
·
m = ρ1v1A1 = u1A1/v1, since density = 1/specific volume
= 50 × 900 × 10–4/0.187
= 24.06 kg/s … Ans. (b)
·
m = ρ2 u2 A2 = u2 A2/v2
·
A2 = m v2/u2
= 24.06 × 0.498/634.43
= 0.01888 ≈ 0.019 m2 … Ans. (c)
Example 2.66 :
A certain water heater operating under steady flow conditions receives 4.2 kg/s of water at
75°C (373 K), of enthalpy 313.93 kJ/kg. The water is heated by mixing with steam. The steam
supplied to the heater is at a temperature of 100.2°C (373.2 K) and is having an enthalpy of
2676 kJ/kg. The mixture leaves the heater as liquid water at a temperature of 100°C (373 K),
having an enthalpy of 419 kJ/kg. Find the steam input rate to the heater in kg/h.
Chemical Engineering Thermodynamics - I 2.147 First Law of Thermodynamics

Solution :
In this case, more than one stream of fluid enters a control surface. For this case, the first law
of thermodynamics, for a steady flow process, gives
2 2 2
·  u1  ·  u2  ·  u3  · ·
m1 H1 + 2 + g z1 + m2 H2 + 2 + g z2 = m3 H3 + 2 + g z3 + Q – Ws … in J/s
     
·
• Assume heater to be well insulated. Therefore, Q = 0.
·
• By the nature of the process it is clear that there is no shaft work. Therefore, Ws = 0.
• Kinetic and potential energy terms are assumed to balance zero.
So the above equation reduces to
· · ·
m1 H1 + m2 H2 = m3 H3 … (1)
[Heat content of input streams = Heat content of output stream]
·
where, m1 = 4.2 kg/s, H1 = 313.93 kJ/kg = 313.93 × 103 J/kg … water at 75°C
·
m2 = ? kg/s, H2 = 2676 kJ/kg = 2676 × 103 J/kg … steam at 100.2°C
· · ·
m3 = m1 + m2 kg/s, H3 = 419 kJ/kg = 419 × 103 J/kg … hot water at 100°C
A mass balance across the heater gives
Input = Output
· · ·
m1 + m2 = m3
· · ·
∴ m3 = 4.2 + m2, m2 = steam input rate in kg/s
Substituting the values in Equation (1), we get
· ·
∴ 4.2 × 313.93 × 103 + m2 × 2676 × 103 = (4.2 + m2) × 419 × 103
· ·
1318.506 + 2676 m2 = 1759.8 + 419 m 2
·
2257 m2 = 441.294
·
∴ m2 = 0.1955 ≈ 0.196 kg/s = 705.6 kg/h
The steam input rate is 705.6 kg/h … Ans.
Example 2.67 :
The following equation connects U, P and V for several gases :
U = a + b PV
where a and b are constants. Prove that for a reversible adiabatic (isentropic) process,
PVγ = constant, where γ = (1 + b)/b
Chemical Engineering Thermodynamics - I 2.148 First Law of Thermodynamics

Solution :
The first law equation, for differential changes, for a closed system is
dU = dQ – dW … differential form
dU = dQ – P dV … for a reversible process
Since the process is adiabatic, dQ = 0
∴ dU = – P dV
dU
dV = – P … (1)

We have, U = a + b PV
Differentiating this equation with respect to V,
dU d (a + b PV) dP
dV = dV = 0 + bV dV + bP

dU dP  dP 
dV = bV dV + bP = b P + V dV … (2)

Equating Equations (1) and (2), we get

–P = b P + V dV
dP
 
dP
–P = bP + bV dV

dP
P (1 + b) + bV dV = 0

dV
Multiplying each term by b PV gives

1 + b dV + dP = 0
 b  V P
dV dP
We know that V = d ln V and P = d ln P

∴ 1 + b d ln V + d ln P = 0
 b 
1+b
Given : b = γ. Therefore,
γ d ln V + d ln P = 0
Integrating,
γ ln V + ln P = constant
ln Vγ + ln P = constant
ln PVγ = constant
∴ PVγ = constant … Ans.
Chemical Engineering Thermodynamics - I 2.149 First Law of Thermodynamics

Example 2.68 :
440 g of air at 180°C (453 K) expands adiabatically to three times its initial volume. During
the process, work done is 52.5 kJ and temperature drops to 15°C. Calculate Cp and Cv.
Solution :
It is a reversible adiabatic expansion process.
Initial temperature = T1 = 180°C (453 K)
Temperature drops to 15°C.
Final temperature = T2 = 15°C (288 K)
Mass of air = m = 440 g = 0.44 kg
Initial volume = V1
Final volume = V2 = 3 V1 (given)
Work done = W = 52.5 kJ
For an adiabatic process,
T2 V1γ–1
= V 
T1  2
288  V1 γ–1
453 = 3V1
0.6357 = (0.333)γ–1
Taking logarithm of both sides,
ln (0.6357) = (γ – 1) ln (0.333)
ln (0.6357) (–0.4530)
γ = 1 + ln (0.333) = 1 + (–1.0996)

γ = 1 + 0.41 = 1.41
Cp
γ = C = 1.41
v

To calculate R :
The work done during a reversible process is given by
mR (T2 – T1)
W = , kJ
γ–1
where R is in J/(kg.K)
W (γ – 1) 52.5 (1.41 – 1)
R = m (T – T ) = 0.44 (453 – 288)
2 1

R = 0.29648 ≈ 0.2965 kJ/(kg.K)


Chemical Engineering Thermodynamics - I 2.150 First Law of Thermodynamics

Cp
We know that : Cp – Cv = R, We have : C = 1.41 ∴ Cp = 1.41 Cv
v
1.41 Cv – Cv = R = 0.2965
∴ Cv = 0.72317 ≈ 0.7232 kJ/(kg.K) … Ans.
Cp – Cv = 0.2965
Cp = 0.2965 + Cv = 0.2965 + 0.7232
∴ Cp = 1.0197 kJ/(kg.K) … Ans.
Example 2.69 :
In a certain steam plant, the power developed by the turbine is 1200 kW. The heat supplied
to steam in the boiler is 3360 kJ/kg. The heat rejected by the system to cooling water in the
condenser is 2520 kJ/kg and the work required by the feed pump to pump the condensate back to
the boiler is 6 kJ/s. Calculate the steam flow rate in kg/s around the cycle.
Solution :

Qin
Boiler Turbine Wout
(+ve)
(+ve)

Feed Qout
pump Condenser
(-ve)

(-ve) Win

Fig. E 2.69
·
Steam flow rate around the cycle = m kg/s
Power developed by the turbine = 1200 kW = 1200 kJ/s
·
Work produced by the turbine = Wout = 1200 kJ/s/m kg/s
·
= 1200/m kJ/kg
Work required by the feed pump = – Win = 6 kJ/s
6 kJ/s 6
= – Win = · = · kJ/kg
m kg/s m
Heat supplied to the boiler = Qin = 3360 kJ/kg
Heat removed from the condenser = – Qout = 2520 kJ/kg
Qout = – 2520 kJ/kg
Chemical Engineering Thermodynamics - I 2.151 First Law of Thermodynamics

Sign conventions for heat and work :


Heat added to the system ⇒ +ve ⇐ work done by the system
Heat removed from the system ⇒ –ve ⇐ work done on the system.


O dQ = algebraic summation of all the heat added round the cycle

= Qin + Qout
= 3360 + (– 2520) = 840 kJ/kg

O dW = Win + Wout =  ·  +  ·  = 1194/m kJ/kg


–6 1200 ·
Similarly, ⌠

m  m 
We have, ⌠
O dQ = ⌠
⌡ ⌡
O dW … first law
·
840 = 1194/m
·
∴ m = 11914/840 = 1.421 kg/s
Steam flow around the cycle = 1.421 kg/s … Ans.
Example 2.70 :
When a system is taken from state 1 to state 2, as shown in Fig. E 2.70, along path 1-3-2
84 kJ of heat flows into the system and it does 32 kJ of work.
(i) How much will be the heat flow into the system along path 1-4-2, if the system does
work of 10.5 kJ along this path ?
(ii) When the system is returned from state 2 to state 1 along the curved path, the work done
on the system amounts to 21 kJ. How much heat is absorbed by or liberated from the
system ?
(iii) Find the heat absorbed in the processes 1-4 and 4-2 if U1 = 0 and U4 = 42 kJ.

3 2

P
1 4

Fig. E 2.70
Solution :
We have, From 1 to 2, along path 1-3-2 :
Heat added to the system = Q1-3-2 = 84 kJ
Work done by the system = W1-3-2 = 32 kJ
Chemical Engineering Thermodynamics - I 2.152 First Law of Thermodynamics

∆U1-2 = Q1-3-2 – W1-3-2 … ∆U = Q – W … first law


U2 – U1 = 84 – 32 = 52 kJ
(i) From 1 to 2, along path 1-4-2 :
Work done by the system = W1-4-2 = 10.5 kJ
Heat added to the system = Q1-4-2 kJ
∆U1-2 = Q1-4-2 – W1-4-2 … from 1 to 2 via 4
(U2 – U1) = Q1-4-2 – 10.5
Q1-4-2 = (U2 – U1) + 10.5
We have, (U2 – U1) = 52 kJ. Therefore,
Q1-4-2 = 52 + 10.5 = 62.5 kJ
Heat that flows into the system = 62.5 kJ … Ans. (i)
(ii) From 2 to 1, along the curved path 2-1 :
Work done on the system = – W2-1 = 21 kJ … by sign convention
i.e., W2-1 = –21 kJ
∆U2-1 = Q2-1 – W2-1
U2 – U1 = Q2-1 + 21
Q2-1 = – (U1 – U2) – 21
= – 52 – 21 = – 73 kJ
The minus sign indicates that the system liberates heat.
Heat liberated by the system (– Q2-1) = 73 kJ … Ans. (ii)
(iii) To find heat absorbed in the processes 1-4 and 4-2 :
Given : U1 = 0 and U4 = 42 kJ
We know that : W1-4-2 = 10.5 kJ
W1-4-2 = W1-4 + W4-2 = 10.5
But W4-2 = 0, since from 4 to 2 volume does not change (no displacement work)
∴ W1-4 + W4-2 = 10.5
W1-4 + 0 = 10.5 ∴ W1-4 = 10.5 kJ
For 1 to 4 : ∆U1-4 = Q1-4 – W1-4
U4 – U1 = Q1-4 – W1-4
Q1-4 = (U4 – U1) + W1-4 = (42 – 0) + 10.5 = 52.5 kJ
Heat absorbed in the process 1-4 = 52.5 kJ … Ans. (iii)
For 4 to 2 : We have,
Q1-4-2 = 62.5 kJ
But, Q1-4-2 = Q1-4 + Q4-2
∴ Q4-2 = Q1-4-2 – Q1-4
= 62.5 – 52.5 = 10 kJ
Heat absorbed in the process 4-2 (Q4-2) = 10 kJ … Ans. (iv)
❐❐❐
Chapter ... 3
P-V-T BEHAVIOUR
AND HEAT EFFECTS
Thermodynamic properties serve to define the state of a system completely (state : each
unique condition of a system). The properties can be classified into two groups : (1) measurable
and (2) nonmeasurable properties. The properties P, V and T are directly measurable, while the
properties U, H, S, etc. are not directly measurable. There exist relations between the
measurable and nonmeasurable properties as well as among the measurable properties P, V
and T.
The P-V-T behaviour of pure fluids can be presented graphically by use of diagrams, in
tabular form or as mathematical relations/expressions between the variables/properties.
The thermodynamic state of a pure fluid can be completely defined by specifying any two of
the three properties - pressure, volume and temperature. The value of the third property depends
upon the values of the other two of the three properties.
Graphical representation of P-V-T behaviour :
The P-V-T behaviour of pure fluids can be represented on the pressure v/s volume diagram
(P-V diagram) and the temperature v/s volume diagram (T-V diagram). In addition to these
diagrams, a pressure v/s temperature - a phase diagram can also be used.
P-V diagram :
Consider the thermodynamic state of a single component system, e.g., water, shown in
Fig. 3.1 as a function of pressure and volume. Fig. 3.1 shows the variation in molar volume (V)
with pressure at different constant temperatures T1, T2, T3 and TC, where T3 > TC > T2 > T1.
On this figure, the liquid and vapour regions are marked as L and V, while the two-phase
region (the area under the dome MCN), where both liquid and vapour phases coexist in
equilibrium, is marked as L + V.

Suppose that the initial state of the system is represented by point A. The change in the
volume of water at a constant temperature T1 occurs along the isotherm A-B-D-N. This isotherm
(the constant temperature line) splits into three segments - AB, BD and DE. The segment AB is
in the liquid region, BD is in the two-phase region and DN is in the vapour region. The slope of
(3.1)
Chemical Engineering Thermodynamics - I 3.2 P-V-T Behaviour and Heat Effects

the isotherm A-B-D-E in the liquid region (the slope of the segment AB of the isotherm ABDE)
is very high (i.e., the isotherm in the liquid region is very steep) because water is almost
incompressible. For incompressible fluids (liquids), liquid volumes change little with large
changes in pressure. Therefore, for the change of state from A to B, a considerable reduction in
pressure results in negligible increase in volume. When point B is reached, the liquid water
begins to boil/vaporise and this action results in a rapid increase in volume. From B to D the
liquid and vapour phases coexist in equilibrium, i.e., they are in equilibrium with each other.
Along the horizontal segment BD, phase transition/change from liquid to vapour takes place at
constant T and P, and more and more of the liquid water converts into the vapour as we move
towards D. Points along the horizontal segment BC represent all possible mixtures of liquid and
vapour in equilibrium, ranging from 100 percent liquid at the point B and 100 percent vapour at
the point D.
The point B represents the saturated liquid state at which the phase change from L to V
starts, while the point D represents the saturated vapour state at which the phase change L to V
ends. The locus of all such points is the dome-shaped curve labelled MCN. The curve MC is
called the saturated liquid curve, while the curve NC is called the saturated vapour curve. At a
given pressure, the temperature at which the phase change from liquid to vapour takes place is
called the saturation temperature. In Fig. 3.1 the saturation temperature is T1 at pressure P1. At a
given temperature, the pressure at which the phase change from liquid to vapour occurs is called
the saturation pressure. At temperature T1, the saturation pressure is P1.
It can be observed that the length of the horizontal segment of the isotherm labelled T1 in the
two-phase region is smaller than the length of the horizontal segment of the isotherm labelled T2,
where T2 > T1. The length of the horizontal segment of any given isotherm in the two-phase
region decreases as the temperature of the system increases. Ultimately, when the temperature of
the system is increased to TC, the horizontal segment reduces to a point (at point C). The point C
is called the critical point and at this point the liquid and vapour phases cannot be distinguished
from each other because their properties are the same. The isotherm passing through the critical
point is called as the critical isotherm. This isotherm exhibits a horizontal inflection at the critical
point C at the top of the dome (MCN).
The temperature and pressure corresponding to the critical point are called as critical
temperature (Tc) and critical pressure (Pc), respectively. The critical properties
(the properties of the system at the critical point) of water are Pc = 220.5 bar, Tc = 647.3 K and
Vc = 56 × 10–6 m3/mol. If the temperature is less than the critical temperature, the substance on
the right of the saturated vapour curve is called a vapour, while the substance is called a gas
if the temperature is above the critical temperature.
Chemical Engineering Thermodynamics - I 3.3 P-V-T Behaviour and Heat Effects

Fig. 3.1 : P-V diagram for a pure fluid (e.g., water)


The system existing at M is a mixture of saturated liquid (state B) and saturated vapour
(state D). As volume is an extensive property, it is additive and therefore it is the molar volume.
The system existing at M is a mixture of saturated liquid (state B) and saturated vapour
(state D). Volume is an extensive property. As an extensive property is additive, the molar
volume V of the mixture at M is equal to the sum of the volumes occupied by the saturated liquid
and saturated vapour.
The quality of a mixture is the ratio of the number of moles of the saturated vapour to the
total number of moles of the mixture. Therefore,
NV mV
x = N + N = m + m where … (3.1)
V L V L

mL – mass of saturated liquid,


mV – mass of saturated vapour,
NL – moles of saturated liquid,
NV – moles of saturated vapour
∴ The molar volume of the mixture at M is
V = xVV + (1 – x) VL … (3.2)
where VV is the molar volume of saturated liquid,
VL is the molar volume of saturated vapour
VLV = VV – VL is the change in molar volume due to vaporisation
Chemical Engineering Thermodynamics - I 3.4 P-V-T Behaviour and Heat Effects

Any extensive property of the mixture can be expressed in terms of the quality of the
mixture. For example, the internal energy of the mixture per mole is given by
U = x UV + (1 – x) UL … (3.3)
where UL is the molar internal energy of the saturated liquid and UV is the molar internal energy
of the saturated vapour.
We have : V = x VV + (1 – x) VL

V = m + m + 1 – m + m  VL
mV VV mV
V L  V L

(mV + mL) V = mV VV + (mV + mL – mV) VL


mV V + mL V = mV VV + mL VL
mL (V – VL) = mV (VV – V)
mV V – VL
∴ mL = VV – V … (3.4)

P-T Diagram/Phase Diagram :


To represent the P-V-T behaviour of a pure substance, in addition to the pressure versus
volume diagram discussed above, a pressure versus temperature plot/diagram or phase diagram
can also be used.
A phase diagram is a plot of temperatures and pressures at which each phase of a pure
substance is the most stable.
A phase diagram is a plot which indicates the conditions of pressure and temperature under
which two or more physical states (phases) of a substance can exist together in a state of dynamic
equilibrium. Therefore, the temperature and pressure are the coordinates of such a diagram.
The diagram (Refer Fig. 3.2) is divided into three single-phase regions and are called solid,
liquid and vapour regions. These regions are separated by three curves. These three curves
represent the conditions of pressure and temperature required for the coexistence of two phases
in equilibrium, i.e., show the conditions of equilibrium between any two of the three phases. The
three curves meet at the triple point where all the three phases-solid, liquid and vapour coexist in
equilibrium. The triple point temperature and pressure of a pure substance is fixed. According to
the phase rule, the triple point is invariant since any change in pressure or temperature will cause
a departure from point O (the triple point), resulting in the disappearance of at least one of the
phases. The system existing along any of these curves (i.e., the two-phase lines) is monovariant
and it is bivariant/divariant in the single-phase regions.
The curve OA is the locus of points where liquid and vapour can coexist in equilibrium. The
curve OA is called the vaporisation curve along which the liquid and vapour phases coexist in
equilibrium.
Chemical Engineering Thermodynamics - I 3.5 P-V-T Behaviour and Heat Effects

The terms saturated liquid and saturated vapour are used to characterise these phases. The
curve OC is called the fusion curve along which the solid and liquid phases coexist in
equilibrium. This curve separates the solid and liquid regions (a plot of the melting point of the
substance at different pressures yields the curve OC, a plot of the vapour pressure of a liquid as a
function of temperature yields the curve OA and a plot of the vapour pressure of a solid as a
function of temperature yields the curve OB). The curve OB is called the sublimation curve
along which the solid and liquid phases coexist in equilibrium and this curve separates the solid
and vapour regions. A single phase is represented by an area or region on the diagram. Thus, area
COA is liquid, COB is solid and BOA is vapour. These one phase areas are divariant in that we
can change two intensive variables, namely pressure and temperature simultaneously and
arbitrarily without causing a new to appear.
The vapour pressure curve/vaporisation curve OA terminates at point A, called the critical
point. The critical temperature - the temperature corresponding to this critical point is the highest
temperature at which a gas can be liquefied by any pressure. The critical pressure is the minimum
pressure needed for the liquefaction of a gas at the critical temperature.

Fig. 3.2 : Phase diagram / P-T diagram for a pure substance


At triple point (point O) : S L V.
OC : liquid vapour – vaporisation curve (liquid-vapour equilibria)
OC : solid liquid – fusion/melting curve (solid-liquid equilibria)
OB : solid vapour – sublimation curve (solid-vapour equilibria)
Mathematical Representation of P-V-T behaviour :
The relationship between the thermodynamic properties such as temperature, pressure and
volume (molar volume) of a single homogeneous fluid is called an equation of state.
Many mathematical expressions have been proposed to represent the P-V-T behaviour of
pure fluids. The relationship of the type
f(P, V, T) = 0
are called equations of state. For any gas, the equation of state is a relation between pressure,
volume and temperature Equation of state gives us a relation between P, V and T when P, V and
Chemical Engineering Thermodynamics - I 3.6 P-V-T Behaviour and Heat Effects

T all vary. The most simple equation of state is the one applicable for ideal gases/perfect gases.
The ideal gas law (or ideal gas equation) is the simplest equation of state. The ideal gas equation
is based partly on the kinetic theory of gases and partly on experiment. The van der Waals
equation is a modification of the ideal gas equation and is derived by means of molecular theory.
Several other equations of state involving many empirical constants are determined from
experimental data, inspite of the fact that their general form may have a theoretical basis.
Ideal gas law :
An ideal gas is the one for which (i) the volume occupied by the molecules themselves is
negligible in comparison with the total volume of the gas and (ii) the intermolecular forces at
attraction between the molecules are negligible.
Thermodynamics characterises an ideal gas as the one for which :
(i) The equation of state is
PV = RT … for one mole of an ideal gas … (3.5)
(ii) The internal energy of an ideal gas depends only on temperature (independent of
P and V) - the internal energy of an ideal gas is a function of temperature only.
∂U = 0 = ∂U and U = f (T)
∴ ∂V  ∂P 
 T  T
∂P
(iii) The Joule-Thomson coefficient (µ) is zero for an ideal gas i.e., µ =   = 0.
∂PH
Boyle's law and Charles' law can be combined to yield the ideal gas equation (ideal gas
equation of state).
Boyle's law states that the molar volume of an ideal gas varies inversely with pressure when
the temperature is held constant.
1
V ∝ P

or PV = constant (at constant T) … (3.6)


For a given mass of an ideal gas the product of the pressure and volume is constant at a
constant temperature.
Charles' law states that the molar volume of an ideal gas is directly proportional to
temperature when the pressure is held constant.
V ∝ T
V
or T = constant (at constant P) … (3.7)
Chemical Engineering Thermodynamics - I 3.7 P-V-T Behaviour and Heat Effects

For a given mass of an ideal gas, the ratio of the volume to temperature is constant at a given
pressure (at a constant pressure).
Combining the above two laws, we get
PV
T = constant … (3.8)

The constant of the above equation is denoted by the symbol R and is called the universal gas
constant. Therefore,
PV = RT … (3.9)
for one mole, where V is the molar volume, i.e., the volume occupied by one mole of the gas,
m3/kmol in SI units. Equation (3.9) is known as the ideal gas law/ideal gas equation for 1 mole of
an ideal gas.
For n moles of an ideal gas, Equation (3.9) becomes
PV = nRT … (3.10)
where V is the volume of n moles.
Equation (3.10) is known as the ideal gas law for n moles of an ideal gas.
The equation of state for n moles of an ideal gas or a perfect gas is given by
PV = nRT
where P is the absolute pressure, V is the volume occupied by n moles and T is the absolute
temperature.
Since one gramole (mol) of an ideal gas occupies a volume of 22.414 l at NTP condition
[T = 273.15 K (0°C) and P = 101.325 kPa (1 atm) - these conditions are said to be normal
temperature and pressure (NTP) conditions].
We have : PV = nRT
PV (1 atm) (22.414 l)
R = nT = (1 mol) (273 K)

= 0.08206 l.atm/(mol.K)

T = 288.7 K (15.6 °C) and P = 101.325 kPa (1 atm) are considered as STP conditions -
standard temperature and pressure.

[In this book, grammoles is abbreviated as mol and kilogrammoles is abbreviated as kmol.]

The value of R for commonly used systems of units is presented in Table 3.1.
Chemical Engineering Thermodynamics - I 3.8 P-V-T Behaviour and Heat Effects

Table 3.1 : Value of universal gas constant (R) in different units


Numerical value of R Units of R = PV/nT
8.31451 m3 · kPa/(kmol·K)
8.31451 J/(mol·K)
0.08206 l · atm/(mol·K)
0.08206 m3 · atm/(kmol·K)
1.987 kcal/(kmol·K)
82.06 cm3 · atm(mol·K)
10.73164 ft3 · psi/(lbmol·°R)
62.36395 cm3 · torr /(mol·K)
0.008314 m3 · MPa/(kmol·K)

The ideal gas law is a relationship of considerable utility. Even though this law does not
apply precisely for any real gas, it is sufficiently accurate for many engineering calculations for
the great majority of gases and vapours at ordinary temperatures and pressures.
The ideal gas law expresses three facts :
(i) the volume of a gas is directly proportional to the number of moles of the gas.
(ii) the volume is directly proportional to the absolute temperature and
(iii) the volume is inversely proportional to the absolute pressure.
For converting gas volumes from one temperature and pressure to another temperature and
pressure, one can use
P1V1 P2V2
T1 = T2


P1  T2
V2 = V1 T  P  … (3.11)
 1  2 
The ideal gas law is valid in the region of low pressure (P → 0) and is only idealisation
towards the actual behaviour of the gases.
Equations of state for real gases :
The behaviour of real gases deviate considerably from that of the ideal gas at elevated
pressures. Thus, this law is inadequate to explain the behaviour of real gases. Many equations of
state have been proposed to describe the P-V-T relationship of real gases. The earliest and the
best known equation of state is that proposed by van der Waals.
Chemical Engineering Thermodynamics - I 3.9 P-V-T Behaviour and Heat Effects

Fig. 3.3 is a graphical representation of Boyle's law. Each curve corresponds to one particular
temperature. The curves are rectangular hyperbola asymptotic to the P-V axis. For any two points
on the curve :
P1 V2
P2 = V1

Fig. 3.3 : P-V relation for an ideal gas at constant T

Fig. 3.4 : T-V relation for an ideal gas at constant P

To derive the equation of state for an ideal gas,

Consider 1 mole of an ideal gas. It changes its state by the following two processes :

(i) Process 1-2' at constant pressure and

(ii) Process 2'-2 at constant temperature.


Chemical Engineering Thermodynamics - I 3.10 P-V-T Behaviour and Heat Effects

Fig. 3.5 : Formulation of equation of state for an ideal gas


At 1 : Pressure = P1, Temperature = T1 and volume = V1

At 2' : Pressure = P'2 = P1 (since P = constant), Temperature = T2' and volume = V2'

At 2 : Pressure = P2, volume = V2 and Temperature = T2 = T2' (since T = constant)


For the process 1-2', applying Charle's law, we have
V1 V2'
T1 = T'
2

But T2 = T2'

V1 V2'
∴ T1 = T2 … (3.12)

For the process 2'-2, applying Boyle's law, we have


P' V' = P V
2 2 2 2

But, we have P'2 = P1

∴ P1V2' = P2V2
P2V2
V'2 = P1 … (3.13)

Substituting the value of V'2 from Equation (3.13) in Equation (3.12) gives
V1 P2V2
T1 = P1T2
P1V1 P2V2
∴ T1 = T2
PV
i.e., T = constant
Chemical Engineering Thermodynamics - I 3.11 P-V-T Behaviour and Heat Effects

The constant in the above equation is taken as R, where R is the universal gas constant.
PV
∴ T = R

PV = RT
which is the ideal gas law or ideal gas equation for 1 mole.
If there are n moles of gas present then the ideal gas equation can be written as
PV = nRT

van der Waals equation of state :


The P-V-T behaviour of real (nonideal) gases differ from that predicted by the ideal gas law,
i.e., real gases considerably deviate from ideal behaviour. In 1873, van der Waals modified the
ideal gas equation (law) so as to make it applicable to real gases.
van der Waals proposed the following equation of state to explain the P-V-T behaviour of
real gases.
P + a  (V – b) = RT … (3.14)
 V 2
for 1 mole of a real gas (V is the molar volume, m3/kmol in SI units), where the constants a and b
are known as the van der Waals constants (they are positive constants). They depend upon the
nature of the gas and can be determined from the experimental P-V-T data. The constant a takes
into account the intermolecular forces of attraction between the molecules and the constant b is
related to the excluded volume of the volume unavailable for molecular motion. When the
constants a and b are zero then the van der Waals equation reduces to the ideal gas equation.
The van der Waals equation is a cubic in volume and accounts for nonideal behaviour of the
gases. For n moles of a real gas, the van der Waals equation is
P + n a (V – nb) = nRT
2
… (3.15)
 V2 
The units of the constants a and b in the SI system are m6.kPa/(kmol)2 and m3/kmol,
respectively wherein V is expressed in m3, P is in kPa and n is in kmol.
The values of the constants a and b can be obtained from the experimental P-V-T data
i.e., by fitting the experimental P-V-T data to the equation of state. These constants can be
estimated in terms of critical constants.
The van der Waals equation being a cubic in V there are three values of V (three roots) for
each pressure at a given temperature, i.e., for a single set of the values of P and T. All of these
roots may be real or two of these roots are imaginary and one is real. A plot of P v/s V at the
Chemical Engineering Thermodynamics - I 3.12 P-V-T Behaviour and Heat Effects

critical temperature of a gas results in a point at which the three values of V become identical.
At this point, the P v/s V curve exhibits a horizontal inflection, so that both the first and second
derivatives of the pressure with respect to the volume, at constant temperature, will be equal to
zero. That is,
 ∂P  = 0  ∂2P  = 0
∂V and ∂V2
 T  T
The van der Waals equation, for one mole of a gas, may be written in the form
RT a
P = V – b – V2 … (3.16)

 ∂P  RT 2a
∂V = – (V – b)2 + V3 …(3.17)
 T
 ∂2P  = 2RT – 6 a
and ∂V2 … (3.18)
 T (V – b)3 V4

At the critical point, the van der Waals equation is given by


P + a  (V – b) = RT
 c V2  c c … (3.19)
 c

and the expression given above [Equations (3.17) and (3.18)] are equal to zero. Therefore,
writing Vc and Tc to indicate the molar critical volume and critical temperature, respectively
(as the volume is critical, V is Vc), we get
–RTc 2a
(V – b)2 + V3 = 0 … (3.20)

2RTc 6a
and (Vc – b)3 – V4 = 0 … (3.21)
c

From Equation (3.21), we have


2RTc 6a
(Vc – b)3 = V4
c

2RTc 3  2a 
(Vc – b) (Vc – b)2 = Vc V3  … (3.22)
 c
From Equation (3.20), we have
RTc 2a
(Vc – b) 2 = 3 … (3.23)
Vc
Chemical Engineering Thermodynamics - I 3.13 P-V-T Behaviour and Heat Effects

2a
Substituting for 3 from Equation (3.23) into Equation (3.22), we get
Vc
2RTc 3  RTc 
(Vc – b) (Vc – b) 2 = Vc (Vc – b)2
2 3
∴ Vc – b = Vc
2Vc = 3Vc – 3b
∴ Vc = 3 b … (3.24)
RTc 2a
We have : (Vc – b) 2 = 3
Vc
Substituting the value of Vc in terms of b from Equation (3.24) in the above equation,
we get
RTc 2a
(3b – b)2 = (3b)3
RTc 2a
4b2 = 27b3
RTc 2a
4 = 27b
8a
∴ Tc = 27Rb … (3.25)
RTc a
We have : Pc = V – b – 2 … (3.26)
c Vc
Substituting the value of Vc from Equation (3.24) and that of Tc from Equation (3.25),
Equation (3.26) becomes
8aR a
Pc = 27Rb (3b – b) – 9b2
8a a 4a a
Pc = – 9b2 = 27b2 – 9b2
2 × 27b2
4a – 3a
Pc = 27b2
a
Pc = 27b2 … (3.27)

Alternatively, we can obtain the van der Waals constants (a and b) in terms of Pc and Tc,
as given below :
2
27R2 Tc RTc
a = 64P and b = 8P
c c
Chemical Engineering Thermodynamics - I 3.14 P-V-T Behaviour and Heat Effects

RTc a
We have : Pc = V – b – 2 but Vc = 3b
c Vc
RTc a
∴ Pc = 3b – b – 9b2
RTc a
Pc = 2b – 9b2

1 RTc a 
Pc = b  2 – 9 b … (3.28)
 
8a a 27RTc
We have : Tc = 27Rb , ∴ b = 8
With this Equation (3.28) becomes
1 RTc 1 27RTc
Pc = b  2 – 9  8 
  
1  c
RT 3RT c 1
Pc = b  2 – 8  = 8 b [4RTc – 3RTc]
 
RTc
Pc = 8 b
RTc
∴ b = 8P … (3.29)
c
8a
We have : Tc = 27Rb
27RTc b RTc
∴ a = 8 , but b = 8Pc … from (3.29)

27RTc RTc
∴ a =  8P 
8  c
2
27R2 Tc
∴ a = 64P … (3.30)
c

Equation (3.14) can be rearranged, as given below :


RT a
P = V – b – V2 … (3.31)

From the above equation it is clear that a plot of P v/s T at constant V gives/yields a straight
line having a slope equal to R/(V – b) and an intercept equal to –a/V2.
Principle of corresponding states :
van der Waals showed that expressing the pressure, temperature and volume of a gas in terms
of its critical pressure, critical temperature and critical volume can result in an important
generalisation called the principle of corresponding states.
Chemical Engineering Thermodynamics - I 3.15 P-V-T Behaviour and Heat Effects

The van der Waals equation of state for one mole of a gas is
P + a  (V – b) = RT … (3.32)
 V 2
where V is the molar volume – i.e., the volume occupied by one mole of the gas – m3/kmol in
SI units.
The ratios of pressure, temperature and volume of a (real) gas to the corresponding critical
values of the gas are called the reduced properties (of the gas).
Pr = P/Pc = reduced pressure
where P is the absolute pressure in question and
Pc is the critical pressure in the same units.
Tr = T/Tc = reduced temperature
Vr = V/Vc = reduced volume, Vc is the molar volume at
the critical point.
We can write : P = PrPc, T = TrTc and V = VrVc
Substituting for P, T and V, Equation (3.32) becomes
P P + 2a 2  (V V – b) = RT T
 r c V V r c r c … (3.33)
 r c
a 8a
We have : Vc = 3b, Pc = 27b2 and Tc = 27b

Substituting for Vc, Pc and Tc, Equation (3.33) becomes

 aPr a  8aTr
27b2 + 2 2 (3bVr – b) = 27b
 9b Vr 

ab  Pr 1  8aTr
2 
b 27 9V + 2  (3V r – 1) = 27b , by taking common a and b
2

 r

 Pr 1  8Tr
∴ 27 + 2 (3Vr – 1) = 27
 9Vr 

Multiplying both the LHS and RHS of the above equation by 27 gives

27Pr 27  27 × 8Tr
 27 + 2 (3Vr – 1) = 27
 9Vr 

P + 3  (3V – 1) = 8T
∴  r V r r r … (3.34)
 
Chemical Engineering Thermodynamics - I 3.16 P-V-T Behaviour and Heat Effects

This equation is known as the reduced equation of state. Equation (3.34) does not involve
either R or the van der Waals constants a and b. Thus, it is a general equation applicable to all
gases (it has general applicability to all gases – generalisation - it is completely general for it
does not involve a and b and hence contains no reference to any specific substance.
It follows from the above equation [Equation (3.34)] that - if two or more gases have the
same reduced pressure and the same reduced temperature then they will have the same reduced
volume. This statement is known as the principle of corresponding states.
OR : At given reduced temperature and reduced pressure, all fluids have the same
reduced volume. The principle/law of corresponding states can be stated mathematically as :
f(Pr, Tr, Vr) = 0.

OR if any two gases (whose P-V-T behaviour can be represented by the van der Waals
equation of state) are at the same reduced temperature and have the same reduced volume then
they must be at the same reduced temperature. The two gases are then said to be in corresponding
states.
Compressibility Charts :
For an ideal gas, PV = RT under all conditions of temperature and pressure, real gases obey
the ideal gas law only approximately and that too under conditions of low pressure and high
temperature.
The deviation of a real gas from ideal gas behaviour, i.e., nonideality of a real gas can be
expressed in terms of the compressibility factor. The compressibility factor Z, which measures
the deviation of a real gas from ideal gas behaviour, is defined as
V PV
Z = (RT/P) = RT … (3.35)

Actual volume of gas


= Volume predicted by ideal gas law

The compressibility factor of a gas is the ratio of the actual volume of the gas to the volume
predicted by the ideal gas law at the same temperature and pressure.
Where P is the pressure of the gas, V is its molar volume and T is the absolute temperature.
The term Z is dimensionless. For an ideal gas Z is unity under all conditions of temperature
and pressure but for a real gas it may be less or greater than unity. Equation (3.35) reduces to the
ideal gas law equation when Z is unity and it can be observed from Fig. 3.6 that, as expected, this
occurs when the pressure approaches zero.
The compressibility factor, Z of any real gas is a function of reduced pressure and reduced
temperature, so that Z = f(Pr, Tr) except near the critical point.
Chemical Engineering Thermodynamics - I 3.17 P-V-T Behaviour and Heat Effects

The compressibility charts are plots of compressibility factor as ordinate against reduced
pressure as abscissa at several constant reduced temperatures.
This chart is an alternative to the use of an equation of state which expresses Z as a function
of Tr and Pr. Inspection shows that Fig. 3.6, i.e., the compressibility chart is a graphical statement
of the general relation/equation f(Tr, Pr, Z) = 0. This chart is constructed by plotting the known
data of one or more gases and can be used for any gas since at a given reduced temperature,
the results of all gases fall on the same curve. Such a chart is shown in Fig. 3.6.

Fig. 3.6 : Generalised compressibility chart

The van der Waals equation of state is

P + a  (V – b) = RT
 V 2
Dividing both sides of the above equation by RT gives
1  a
RT P + V2 (V – b) = 1

Multiplying both sides by V gives


V  a
RT P + V2 (V – b) = V

PV + aV  (V – b) = V
RT RTV2
Chemical Engineering Thermodynamics - I 3.18 P-V-T Behaviour and Heat Effects

PV
We have : RT = Z = compressibility factor
Z + a  (V – b) = V … (3.36)
 VRT
2
27R2 Tc RTc
We have : a = 64P and b = 8P
c c

Substituting the values of a and b, Equation (3.36) becomes


 27R2 Tc  
2

 V – c = V
RT
Z +
 64Pc VRT  8Pc 

 27RTc  
2

 V – c = V
RT
Z +
 64VTPc  8Pc 
We have : Tr = T/Tc ∴ Tc = T/Tr, Pr = P/Pc ∴ Pc = P/Pr
Substituting for Tc and Pc, the above equation becomes

 27RT2Pr   RTPr
2  V –
Z +
64 VT Tr P  8TrP  = V

 27Pr RT  RTPr
Z + 2  V – 8PTr  = V … (3.37)
 64Tr PV 
PV RT 1 RT V
We have : Z = RT ∴ PV = Z and P = Z

With this, Equation (3.37) becomes


 27Pr   VPr 
2  V –
Z +
64Z Tr   8ZTr = V

Taking V common from the terms in the second bracket on the LHS, we get
 27Pr   Pr 
Z + 2 V 1 – 8ZT  = V
 64Z T r r

 27Pr   Pr 
Z + 2 1 –  = 1 … (3.38)
 64Z Tr   8Z Tr
Equation (3.38) shows that all gases when considered at given values of Tr and Pr will have
the same value of compressibility factor. The compressibility of any gas is a function of the
reduced temperature and reduced pressure. The principle of corresponding states can be stated
mathematically as :
Z = f(Tr, Pr)
Chemical Engineering Thermodynamics - I 3.19 P-V-T Behaviour and Heat Effects

If two or more gases have the same reduced temperature and the same reduced pressure then
they will have the same compressibility factor and they deviate from the ideal behaviour to the
same extent. This statement is known as the principle of corresponding states.
Two or more gases having the same reduced temperature and the same reduced pressure
(i.e., having the same values of Tr and Pr) and thus having the same reduced volume are said to
be in corresponding states.
Evaluation of van der Waals constants for a gas mixture :
To evaluate a and b for a gas mixture we have to evaluate Tc and Pc of the gas mixture as per
the procedure given below :
According to Kay's rule the gas mixture, can be treated as if it were a pure single-component
gas with a pseudo-critical temperature and pressure.
Pseudo-critical properties of components present in the gas mixture are evaluated using Kay's
rule.
Pseudo-critical property
of a component in = (mole fraction of the component) ×
the gas mixture
(critical property of the component)
For a gas mixture containing A, B and C, we have :
Pseudo-critical
temperature of A = xA Tc, A
where, xA = mole fraction of A and Tc, A = critical temperature of A
Pseudo-critical temperature of B = xB Tc, B
Pseudo-critical temperature of C = xC Tc, C
The pseudo-critical temperature of the gas mixture according to Kay's additive rule is
Tc, mix = xA Tc, A + xB Tc, B + xC Tc, C … (3.39)
Similarly, Pc, mix = xA Pc, A + xB Pc, B + xC Pc, C … (3.40)
where, Pc, A = critical pressure of pure A, etc.
Pc, mix = pseudo-critical pressure of the gas mixture
Evaluate a and b using Tc, mix and Pc, mix.
Virial Equation of State :
The virial (a Latin word used for force which refers to interaction forces between molecules)
equation of state, suggested/proposed by Kammerlingh Onnes for real gases, wherein the
Chemical Engineering Thermodynamics - I 3.20 P-V-T Behaviour and Heat Effects

compressibility factor (Z) of a gas or vapour is expressed as a power series in reciprocal volumes
(1/V) is
PV B C D
Z = RT = 1 + V + V3 + V3 + … … (3.41)

where V is the molar volume (m3/kmol) and B, C and D are known as the second, third and
fourth virial coefficients, respectively. For a given gas, these coefficients are functions of
temperature only, i.e., they depend on temperature and type of species. The virial coefficients can
be determined from experimental P-V-T data. The second virial coefficient takes into account the
deviation from ideal behaviour which results from the interactions of two molecules and the third
virial coefficient takes into account the deviation from ideal behaviour which results from the
interactions of three molecules and so on.

This equation is explicit in pressure [often called the Leiden form or pressure explicit
(i.e., volume and temperature appear as independent variables) form of the virial equation] and
can be made to represent the experimental P-V-T data more accurately by increasing the number
of terms but it can be truncated after the second term without much error to represent the P-V-T
data.

The compressibility factor representing the deviation from ideal behaviour (i.e., non-ideality
of a gas) is defined as
PV PV V V
Z = (PV) = RT = (RT/P) = (V)
ideal ideal

Volume of real gas


= Volume of the gas behave ideally at the same T and P
(i.e.‚ volume predicted by ideal gas law)

where V is the molar volume (m3/kmol), i.e., the volume occupied by one mole of the gas.

The compressibility factor of any real gas is the ratio of the volume of the real gas to the
volume if the same gas behaves ideally at the same temperature and pressure. For an ideal gas
Z = 1.0 under all conditions of temperature and pressure. The value of Z for any real gas may be
less than or greater than unity depending upon P and T of the gas. Z is a measure of the
nonidealness of the gas under consideration.

The virial equation can also be written as a power series in pressure, as given below :
PV
RT = 1 + B'P + C'P + D'P + …
2 3 … (3.42)

This equation is explicit in volume. It is called the Berlin form or the volume explicit virial
equation of state.
Chemical Engineering Thermodynamics - I 3.21 P-V-T Behaviour and Heat Effects

Where B', C' and D' are the second, third and fourth virial coefficients. They depend on
temperature and type of species. The two sets of virial coefficients are related to each other,
as given below :
B
B' = RT , C' = (C – B2)/(RT)2

D – 3BC + 2B2
and D' = (RT)3 … (3.43)

These values of B', C' and D' are obtained in Ex. 3.30. Please refer to Ex. 3.30.
Redlich-Kwong Equation :
The Redlich-Kwong equation is a comparatively simple two-parameter (involving two
constants) equation of state. The RK equation of state is given by
RT a
P = V – b – T1/2 V (V + b) … (3.44)

where a and b are empirical constants. The values of the constants a and b can be evaluated using
the following equations :
2.5
0.4278 R2 Tc
a = Pc … (3.45)

0.08664R Tc
and b = Pc … (3.46)

In Equation (3.44), V is the molar volume (m3/kmol).


The R-K equation is the best two-parameter equation of state for real gases. It is in wide use
for expressing the P-V-T behaviour of gases due to its simplicity and accuracy.
Soave-Redlich-Kwong Equation of State :
Soave proposed a modification to the Redlich-Kwong equation to improve its accuracy in the
predicted data by introducing a third parameter (α). The Soave-Redlich-Kwong (SRK) equation
of state is given by
RT aα
P = V – b – V(V + b) … (3.47)
2
0.42748 R2 Tc
where, a = Pc
0.08664 RTc
b = Pc
1/2
α = [1 + (0.480 + 1.57 ω – 0.176 ω2) (1 – Tr )]2
where ω is the ascentric factor.
This equation is widely used to predict the properties of hydrocarbons.
Chemical Engineering Thermodynamics - I 3.22 P-V-T Behaviour and Heat Effects

Peng-Robinson (PR) equation of state :


The PR equation of state is given by
RT aα
P = V – b – V(V + b) + b(V – b) … (3.48)
2
0.45724 R2 Tc
where, a = Pc
RTc
b = 0.07780 P
c
1/2
and α = [1 + (0.37464 + 1.542 ω – 0.26992 ω2) (1 – Tr )]2
This equation is used to predict the properties of hydrocarbons and inorganic gases such as
N2, O2, etc.
HEAT EFFECTS

Heat transfer constitutes an integral and very often inevitable part of every chemical process.

The heat effects - transfer of heat or temperature changes, or both are associated with
physical processes as well as with chemical reactions.

In the phase change operation of a pure substance from liquid to vapour at constant pressure,
there is no change in temperature but there is a definite transfer of heat from the surroundings to
the substance. This heat effect is called the latent heat of vaporisation.

In this section, we will deal with the heat effects associated with chemical reactions
[–transfer of heat or temperature changes during the course of reaction, or both).

Usually, chemical reactions are accompanied by energy absorption or rejection as heat,


i.e., when chemical reactions take place, heat is either absorbed or evolved. This is due to the fact
that the energy associated with the products, i.e., the heat content of the products is different
from the energy associated with the reactants, i.e., the heat content of the reactants and this
difference in the heat contents arises because of the differences in molecular structure (the
difference in the structure of molecules of the reactants and products).

Reactions which are accompanied by evolution of heat are called exothermic reactions.
In these reactions, Hp < HR so that ∆H is negative [HP is the heat content/enthalpy of the products
and HR is the enthalpy of the reactants.]

Reactions which are accompanied by the absorption of heat (reactions to which heat must be
added to maintain constant temperature) are called endothermic reactions. In these reactions,
HP > HR, so that ∆H is positive.
Chemical Engineering Thermodynamics - I 3.23 P-V-T Behaviour and Heat Effects

(i) Exothermic reactions (ii) Endothermic reactions


Fig. 3.7 : Energy level diagram for chemical reactions
Standard states :
Chemical reactions are invariably associated by enthalpy changes. They can be carried out at
different conditions of temperature and pressure. Since it is not possible to tabulate the enthalpy
changes at all possible conditions of temperature and pressure, it is necessary to standardise the
chemical reactions. For this purpose, the following standard states of chemical specifies at a
given temperature are widely accepted.
Gases : Pure gas in the ideal gas state at 1 bar (0.1 MPa)
Liquids : Pure liquid at 1 bar (0.1 MPa)
Solids : Pure solid at 1 bar (0.1 MPa)
(A standard state is the particular state of a species at temperature T defined by generally
accepted conditions of pressure, composition and physical state (or state of aggregation) of the
species. The standard state pressure - the reference pressure for the standard state - is 1 bar
(0.10 MPa). The physical state for gases is the ideal-gas state and for liquids and solids it is the
real liquid state and the real solid state respectively at the system temperature (T) and at 1 bar.
The symbols commonly used to indicate the gaseous liquid and solid states (of the species
involved) are g, l and s respectively. In thermochemical equations, the symbol g, l or s is placed
in parentheses after the chemical formula to make clear whether the species is a gas, a liquid or
a solid. For example,
C (s) + O2 (g) → CO2 (g) … ∆H = –393509 J
The standard states with respect to composition are states of the pure species, i.e., the species
are assumed to be pure components in the standard state.
Heat of Reaction :
• The heat change accompanying a chemical reaction is equal to the difference between
the total heat content of the products and that of the reactants at constant temperature
and pressure and this is referred to as the heat of reaction.
• The enthalpy change accompanying a chemical reaction at constant temperature is
called as the heat of reaction. It is the difference between the enthalpy of the products
and that of the reactants. It is denoted by the symbol ∆H.
Chemical Engineering Thermodynamics - I 3.24 P-V-T Behaviour and Heat Effects

∆H values may be positive or negative. If ∆H is positive, i.e., the enthalpy (or heat content)
of the products is more than that of the reactants, the reaction is accompanied by an absorption of
heat (the reaction is called endothermic). On the other hand if ∆H is negative, i.e., the enthalpy of
the reactants is more than that of the products, the reaction is accompanied by an evolution of
heat (the reaction is accompanied by decrease in enthalpy (and is called exothermic reaction).
Exothermic reactions take place with decrease in enthalpy, while endothermic reactions take
place with increase in enthalpy.

Standard heat of reaction :

The enthalpy change associated with a chemical reaction wherein the reactants and the
products are in their respective standard states at the same temperature T is called as the standard
heat of reaction (at temperature T). It is denoted by the symbol ∆H° , where the superscript o
T

(zero) indicates the standard state, i.e., it indicates that all the reactants and products are in their
standard states and the subscript T indicates the temperature at which the reaction is carried out.
It is also called the standard enthalpy change of the reaction.

Whenever a heat of reaction is given for a particular reaction, it is applicable for the
stoichiometric coefficients as written. If each of the stoichiometric coefficient is doubled then the
heat of reaction is doubled. For example, the methanol synthesis reaction may be written as
1 1 ° = –45.07 kJ
(A) 2 CO (g) + H2 (g) → 2 CH3OH (g), ∆H298

or (B) CO (g) + 2 H2 (g) → CH3OH (g), ° = –90.14 kJ


∆H298

The stoichiometric coefficients in Equation (A) are multiplied by a factor 2 in rewriting the
Equation (B). Therefore, ∆H° for reaction (B) is equal to twice that for the reaction (A).
298

° indicates that the heat of reaction is the standard value for a temperature
The symbol ∆H298
of 298.15 K (25 °C). In thermodynamics, by conventions the standard heats of reaction are
reported at a temperature of 298.15 K (25 °C).

Standard heat of formation :

A reaction in which only one mole of a compound is formed from its constituent elements is
called a formation reaction. For example, the reaction
1
C(s) + 2 H2 (g) + 2 O2 (g) → CH3OH (g)

is the formation reaction for methanol.


Chemical Engineering Thermodynamics - I 3.25 P-V-T Behaviour and Heat Effects

The enthalpy change accompanying the formation of one mole of a compound from its
constituent elements is called the enthalpy of formation or heat of formation. As formation
reactions result in the formation of 1 mole of the compound, the heat of formation is based on
1 mole of the compound formed (Here 1 mole is 1 grammole, abbreviated as 1 mol). It is denoted
by the symbol ∆Hf, where the subscript f indicates that it is the heat of formation.

The heat of formation of a compound is positive when the formation of the compound from
its constituent elements accompanies with increase in enthalpy, while it is negative when the
formation reaction accompanies with decrease in enthalpy. A compound with negative heat of
formation is called as exothermic compound, while a compound with positive heat of formation
is called as an endothermic compound.
The enthalpy change associated with a formation reaction, when the reactants and products
are in their standard states is called the standard heat of formation.
The enthalpy change associated with the formation of 1 mole of a compound from its
elements when all the species of the reaction are in their standard states is called the standard
heat of formation (or the standard enthalpy of formation of the compound) : By convention, the
standard heats of formation of elements are taken as zero at all temperatures. The heat of
formation of a compound is the difference between the heat content (enthalpy) of the compound
and that of its elements. Since the latter are taken as zero, it follows on the basis of this
convention that the heat content of a compound is equal to its heat of formation. The standard
°
heat of formation is denoted by the symbol ∆Hf where the superscript o (zero) indicates the
standard state/indicates that it is the standard value.
For tabulating the data, the standard heats of formation of compounds are reported at
298.15 K (25 °C). The standard heat of formation of a compound at a temperature of 298.15 K is
°
represented by the symbol ∆Hf 298 , where the superscript o (zero) indicates the standard state and
the 298 is the approximate absolute temperature in kelvins (K).
Standard heat of combustion :
A combustion reaction is defined as the reaction between an element or a compound and
oxygen to form specific products. In case of combustion of organic compounds containing
carbon, hydrogen and oxygen only, the products of combustion are carbon dioxide and water (the
physical state of water may be liquid or vapour, i.e., it may be present as liquid or vapour).
The enthalpy change accompanying the complete combustion of 1 mole (1 grammole,
abbreviated as 1 mol) of a compound with molecular oxygen is called the enthalpy of combustion
or heat of combustion. It is denoted by the symbol ∆Hc.
Chemical Engineering Thermodynamics - I 3.26 P-V-T Behaviour and Heat Effects

In reality only a few formation reactions can be actually carried out and therefore the heat of
formation is usually determined indirectly. A reaction such as the formation of n-butane :
4C (s) + 5 H2 (g) → C4H10 (g) is not practically feasible and therefore, it is not possible to
directly measure the standard heat of formation of C4H10 (g). However, the combustion of
n-butane can be carried out in a flow calorimeter to determine the enthalpy change. Standard
heats of formation which cannot be determined directly can be calculated from standard heats of
combustion.
The standard heat of combustion of a substance at temperature T is the enthalpy change
accompanying the complete combustion of 1 mole of the substance with molecular oxygen when
all the reactants and products are in their standard states and at the same temperature T. It is
°
denoted by the symbol ∆Hc , where the superscript o (zero) indicates the standard state and the
subscript c indicates that it is the heat of combustion (or the subscript c denotes the combustion
reaction). For convenience of tabulating the data, the standard heats of combustion of compounds
°
are reported at 298.15 K (25 °C) and are represented by the symbol ∆Hc 298 reported in kJ/mol.

Standard heat of reaction from standard heats of formation :


The standard heat of any reaction, i.e., the standard enthalpy change of any reaction at
298.15 K (25 °C), ∆H°298 , can be calculated from a knowledge of the standard heats of
formation/standard enthalpies of formation of the species taking part in the reaction.
The standard enthalpy change of a reaction (the standard heat of reaction) is equal to the
difference of the sum of the standard heats of formation of the products and the sum of the
standard heats of formation of the reactants, i.e.,

∆H°298 = Enthalpy of products – Enthalpy of reactants

° °
= ∑ ∆Hf 298 (products) – ∑ ∆Hf 298 (reactants) … (3.49)

Consider the reaction between a moles of A and b moles of B to produce r moles of R and
s moles of S
aA + bB → rR + sS
° °
∆H°298 = ∑ ∆Ηf 298 (products) – ∑ ∆Hf 298 (reactants)

° ° ° °
= [r ∆Hf 298‚ R + s ∆Hf 298‚ S ] – [a ∆Hf 298‚ A + b ∆Hf 298‚ B ] … (3.50)

when ∆H° is positive, the reaction is said to be endothermic, i.e., it absorbs heat during the
course of reaction and when ∆H° is negative, the reaction is said to be exothermic, i.e., heat is
evolved during the course of reaction (i.e., while the reaction is taking place).
Chemical Engineering Thermodynamics - I 3.27 P-V-T Behaviour and Heat Effects

If a chemical reaction is represented by the equation ∑ viAi = 0, where vi represents the


stoichiometric coefficient associated with the species Ai involved in the reaction, then the sign
convention used for the stoichiometric coefficients is that the stoichiometric coefficient
associated with a product is positive while that associated with a reactant is negative. Therefore,
the standard heat of reaction at 298.15 K is given by
°
∆H°298 = ∑ vi ∆Hf 298 i … (3.51)

°
where ∆Hf 298‚ i is the standard heat of formation at 298.15 K (25°C) of the ith component/species.

Standard heat of reaction from standard heats of combustion :


The standard enthalpy change of a reaction at 298.15 K (the standard heat of reaction)
involving organic compounds can be calculated from a knowledge of the standard enthalpies of
combustion of the species at 298.15 K (25 °C) taking part in the reaction.
The standard heat of reaction at 298.15 K (25 °C) is equal to the difference between the sum
of the standard heats of combustion (the standard enthalpies of combustion) of the reactants and
the sum of the standard heats of combustion of the products (at 298.15 K).
° °
∆H°298 = ∑ ∆Hc 298 (reactants) – ∑ ∆Hc 298 (products) … (3.52)

For the reaction aA + bB → rR + sS, we can write


° ° ° °
∆H°298 = [a ∆Hc 298‚ A + b ∆Hc 298‚ B ] – [r ∆Hc 298‚ R + s ∆Hc 298‚ S ] … (3.53)

If a chemical reaction is represented by the equation ∑ vi Ai (where vi represents the


stoichiometric coefficient of the species Ai, Ai is the chemical formula of species i, vi is positive
for products and negative for reactants), the standard heat of reaction is given by
°
∆H°298 = – ∑ vi ∆Hc 298‚ i … (3.54)

°
where ∆Hc 298‚ i is the standard heat of combustion of species i.

[In some books for tabulating the data of standard heats of formation and standard heats of
combustion, the standard state temperature is taken as 298.15 K (25 °C). With this, the standard
states are - gas : pure in the ideal gas state at 298.15 K and 1 bar (0.1 MPa), liquid : pure at
298.15 K and 1 bar and solid : pure crystalline at 298.15 K and 1 bar. In these books, the
standard heat of formation and standard heat of combustion at 298.15 K (25 °C) are represented
° ° ° °
by ∆Hf and ∆Hc respectively (instead of representing them by ∆Hf 298 and ∆Hc 298 ). Similarly,
Chemical Engineering Thermodynamics - I 3.28 P-V-T Behaviour and Heat Effects

the standard heat of reaction at 298.15 K is represented by ∆HR° instead of ∆H298


° . With these

symbols, we have
Standard heat of reaction at 298.15 K
° °
= ∆H°R = ∑ ∆Hf (products) – ∑ ∆Hf (reactants)
° °
∆H°R = ∑ ∆Hc (reactants) – ∑ ∆Hc (products) ]

Hess's law of constant heat summation :


It states that the net enthalpy change (i.e., the net heat evolved or absorbed) in a given
chemical reaction is the same whether the reaction takes place in one step or in several steps
(i.e., in a series of steps). This means that the net enthalpy change, i.e., the net heat of reaction
depends on the pressure, temperature, state of aggregation, initial and final states and is
independent of the number of intermediate states through which the reaction system may have
passed, i.e., independent of the manner in which the reaction takes place.

This law permits us to treat thermochemical equations as algebraic equations –


thermochemical equations can be added, substracted and multiplied just like algebraic equations
- the enthalpies of reactions can be added, substracted and multiplied algebraically. Using this
law, we can calculate the heat of formation of a compound from a series of reactions that do not
involve the direct formation of the compound from its constituent elements. The heat of
formation of a compound can be calculated if the heat of combustion data of all the species
taking part in the formation reaction are known. For example,

According to Hess's law,


∆H = ∆H1 + ∆H2 + ∆H3 + ∆H4

Hess's law is useful to find out the enthalpy changes of many reactions and enthalpies of
formation of many compounds which cannot be determined experimentally. For example, it is
very difficult to measure the heat evolved during the combustion of carbon to carbon monoxide.
Chemical Engineering Thermodynamics - I 3.29 P-V-T Behaviour and Heat Effects

1
(3) C(s) + 2 O2 (g) → CO (g) ∆H3° = ?

However, we can measure the heat evolved during

(1) C(s) + O2 (g) → CO2 (g) ∆H1° = – 393.5 kJ/mol … at 298 K


1
(2) CO (g) + 2 O2 (g) → CO2 (g) ∆H2° = –283 kJ/mol … at 298 K

Substracting (2) from (1) gives (3)


1
C(s) + 2 O2 (g) → CO (g) … (3)

Therefore, ∆H°3 = ∆H1° – ∆H2° = –393.5 – (–283) = –110.5 kJ/mol

It is impossible to determine experimentally the enthalpy of formation of benzene from its


elements. But, the same can be calculated from the known heat of combustion of benzene and
enthalpies of formation of CO2 and H2O (l).
The thermochemical equations to evaluate the heat of formation of benzene are :

(i) C6H6 (l) + 7.5 O2 (g) → 6 CO2 (g) + 3 H2O (l) ∆H1° = –3267.7 kJ/mol … at 298 K

(ii) C(s) + O2 (g) → CO2 (g) ∆H2° = –393.5 kJ/mol


1
(iii) H2 (g) + 2 O2 (g) → H2O (l) ∆H3° = –285.8 kJ/mol

(iv) 6 C(s) + 3 H2 (g) → C6H6 (l) ∆H4° = ?

Obtain Equation (iv) from Equations (i), (ii) and (iii).


Equation (iv) = 6 Equation (ii) + 3 Equation (ii) – Equation (i)
= 6C(s) + 6 O2 (g) → 6 CO2 (g)
+ 3 H2 (g) + 1.5 O2 (g) → 3 H2O (l)
– [C6H6 (l) + 7.5 O2 (g) → 6 CO2 (g) + 3 H2O (l)]
= 6C (s) + 3 H2 (g) → C6H6 (l)

Therefore,
°
∆H°4 = ∆Hf C = 6 ∆H2° + 3 ∆H3° – ∆H1°
6H6

= 6 (–393.5) + 3 (–285.8) – (–3267.7)


= 49.3 kJ/mol … at 298 K
Chemical Engineering Thermodynamics - I 3.30 P-V-T Behaviour and Heat Effects

Effect of temperature on standard heat of a reaction :


Consider the reaction
aA + bB → rR + sS
For this reaction, the standard heat of reaction (the standard enthalpy change for the reaction)
is given by

∆H° = 
Standard-state enthalpy Standard-state enthalpy
of products  – of reactants 
°
or ∆H° = ∑ vi Hi … (3.55)
°
where Hi is the enthalpy of species i in its standard state and the summation is over all products
and reactants.
∆H° = (rHR° + sHS° ) – (aHA° + bHB° ) … (3.56)
Differentiating this equation with temperature, keeping pressure constant, we get
 °  °  °  °
∂(∆H°) = r ∂HR + s ∂HS  – a ∂HA – b ∂HB
 ∂T  … (3.57)
 P  ∂T P  ∂T P  ∂T P  ∂T P
We know that Cp = (∂H/∂T)P … (3.58)
i.e., the heat capacity of a system at constant pressure is equal to the rate of the increase of
enthalpy (heat content) of the system with temperature at constant pressure. CPi is the molar heat
capacity of species i.
∂(∆H°) = r C° + s C° – a C° – b C°
∴  ∂T  … (3.59)
 P PR PS PA PB

Let ∆Cp = Sum of heat capacities of products – Sum of heat capacities of


reactants at a given T.
∂(∆H°) = ∆C°
 ∂T  … (3.60)
  p

°
∆Cp is the increase in heat capacity of the system accompanying the chemical reaction.

Equation (3.60) is called as the Kirchhoff equation. It states that the rate of variation of the
heat of reaction with temperature, at constant pressure, is equal to the increase of the heat
capacity of the system accompanying the reaction.
Equation (3.60) can be written as
°
d (∆H°) = ∆Cp dT … (3.61)

This equation relates heats of reaction to temperature.


Chemical Engineering Thermodynamics - I 3.31 P-V-T Behaviour and Heat Effects

The Kirchhoff equation is used to determine the standard heat of reaction at one temperature
if we know it at another temperature. This involves integration of the above equation between the
temperature limits of T1 and T2.
T2 T2
°

⌡ d(∆H°) = ⌠
⌡ ∆Cp dT … (3.62)
T1 T1
T2
° ° °
⌠ ∆Cp dT
∆HT2 – ∆HT1 = ⌡ … (3.63)
T1

° °
where ∆HT2 – ∆HT1 are the standard heats of reaction at the temperatures T1 and T2.

If the limits are 298.15 K and T, then


298 T

⌡ d(∆H°) = ⌠
⌡ ∆Cp dT
T 298
T
∆H°T – ∆H298
° = ⌠ ∆C dT
⌡ p … (3.64)
298

If the temperature range is small, the heat capacities are assumed to be independent of
temperature. ∆Cp is independent of temperature and it can be taken outside the integral sign,
which gives
° ° °
∆HT2 – ∆HT1 = ∆Cp (T2 – T1) … (3.65)
i.e., ∆H° changes linearly with increase of temperature.
° °
If ∆Cp is taken as the mean value ∆Cpm in the temperature range from T1 to T2, then
° ° °
∆HT2 – ∆HT1 = ∆Cpm (T2 – T1) … (3.66)
°
where Cpmi – mean molar heat capacity of species i
If the temperature range is not small, then the constancy of the heat capacities is not a valid
assumption and we must express the heat capacities as a function of temperature. The variation
of the heat capacity with temperature of many substances can be expressed as a power series in
T, viz.,
°
Cp = α + βT + γT2 + …

where α, β and γ are constants for a given species. Therefore,


°
∆Cp = [(rαR + sαS) – (aαA + bαB)] + [(rβ R + sβ S) – (aβ A + bβ B)] T +
+ [(rγR + sγS) – (aγA + bγB)] T2 + …
°
∆Cp = ∆α + ∆βT + ∆γT2 + … … (3.67)
Chemical Engineering Thermodynamics - I 3.32 P-V-T Behaviour and Heat Effects

Inserting Equation (3.67) into Equation (3.63) and carrying out integration gives
° ° ∆β 2 2 ∆γ 3 3
∆HT2 – ∆HT1 = ∆α (T2 – T1) + 2 (T2 – T1) + 3 (T2 – T1) + … … (3.68)

Equation (3.68) is the integrated Kirchhoff equation.


So the standard heat of reaction at one temperature can be calculated from the known
standard heat of reaction at another temperature provided heat capacity data are available for all
the species involved.
If is often useful to derive a general equation which will give the standard enthalpy
change/standard heat of a reaction at any temperature T.
Consider the reaction
aA + bB→ rR + sS
For this reaction we have to determine the standard heat of reaction at temperature T, ∆HT° if
we know the standard heat of reaction at a temperature of 298.15 K. Since the standard heat of
reaction is independent of the path by which the reactants are converted in products, we can
replace the actual path or process by a series of convenient imaginary paths for the determination
of ∆H° .
T

The standard heat of reaction for the reaction at temperature T (T > 298.15 K) can be
determined by following the three-step path or process shown in Fig. 3.8.

Fig. 3.8 : Effect of temperature on heat of reaction


(a three-step process connecting the reactants at T to product at T)
Step 1 : The reactants in their standard states are cooled from temperature T to 298.15 K.
The enthalpy change for this step is given by
298
°
∆H1° = ∑ ⌠
⌡ ni Cpi dT … (3.69)
reactants T

°
where the summation is over all the reactants, ni is the moles and Cpi is the molar heat capacity
of species i.
Chemical Engineering Thermodynamics - I 3.33 P-V-T Behaviour and Heat Effects

Step 2 : The reactants in their standard states at 298.15 K are converted into products in
their standard states at 298.15 K, i.e., the reaction is allowed to take place at 298.15 K. The
enthalpy change associated with this reaction step at 298.15 K is ∆H° . 298

°
∆H°2 = ∆H298 … (3.70)

Step 3 : In this step, the temperature of the products in their standard states is raised from
298.15 K to T. The enthalpy change associated with this step is given by
T
∆H3° = ∑ ⌠
⌡ ni Cpi dT … (3.71)
products 298

The standard heat of a reaction at temperature T is obtained by adding the preceding three
equations since the enthalpy change is independent of path.

[for all practical purposes, T = 298.15 K ⇒ 298 K]

Actual path : Reactants at T to products at T.

Devised path : Reactants at T → reactants at 298.15 K → reactants at 298.15 K → products


at T.
Step 1 to 4 = step 1 to 2 + step 2 to 3 + step 3 to 4

∆H°T = ∆H1° + ∆H2° + ∆H3° … (3.72)

° + ∆H° since ∆H° = ∆H°


= ∆H1 + ∆H298 3 2 298

298 T
° ° °
= ∑ ⌠
⌡ ni Cpi dT + ∆H298 + ∑ ⌠
⌡ ni Cpi dT
reactants T products 298

T T
° °
= ∆H°298 + ∑ ⌠
⌡ ni Cpi dT – ∑ ⌠
⌡ ni Cpi dT
products 298 reactants 298

… (3.73)
° ° °
We have : ∑ (nCp )products – ∑ (nCp )reactants = ∆Cp .

It is the increase of the heat capacity of the system accompanying the chemical reaction,
where n is the number of moles of the substance taking part in the reaction (the respective
°
stoichiometric coefficients of the species involved in the reaction) and Cp is the molar heat
°
capacity. ∆Cp is the difference between the sum of heat capacities of the products and the
reactants at a given temperature.
Chemical Engineering Thermodynamics - I 3.34 P-V-T Behaviour and Heat Effects

° °
∑ (n Cp )products = ∑ (ni Cpi ) and the same is true for the reactants.
products
T
∴ ° + ⌠ ∆C° dT
∆H°T = ∆H298 ⌡ p … (3.74)
298

If the temperature dependence of the molar heat capacity of the species involved is given by,
°
Cp = α + βT + γT2 + …

where α, β, γ … are constants, and T is the absolute temperature, then it follows that
°
∆Cp = ∆α + ∆βT + ∆γT2 + …

where, ∆α = (r αR + s αS) – (a αA + b αB)


∆β = (r β R + s β S) – (a β A + b β B)
and ∆γ = (r γR + s γS) – (a γA + b γB)

OR for aA + bB + …… → rR + sS + …
∆α = (r αR + s αS + …) – (a αA + b αB + …)

∆β = (r β R + s β S + …) – (a β A + b β B + …)

∆γ = (r γR + s γS + …) – (a γA + b β B + …)

Equation (3.74) can be rewritten as


T
° + ⌠ (∆α + ∆βT + ∆γT2 + …) dT
∆HT° = ∆H298 ⌡
298

° + ∆α + (T – 298) + ∆β (T2 – 298


∆HT° = ∆H298
2
)
2

∆γ 3
+ 3 (T3 – 298 ) + … … (3.75)

Grouping the constants appearing in the above equation into a single constant ∆Ho, we get

∆β ∆γ
∆H°T = ∆Ho + ∆αT + 2 T2 + 3 T3 + … … (3.76)

The constant ∆Ho is evaluated from a knowledge of ∆HT° at any temperature. The value of
° . Put ∆H° = ∆H°
∆Ho is evaluated from the standard heat of reaction at 298.15 K, i.e., ∆H298 T 298
value and T = 298 K to get ∆Ho.
Chemical Engineering Thermodynamics - I 3.35 P-V-T Behaviour and Heat Effects

°
If Cp is expressed as
°
Cp = α + βT + γT2 + δT–2 , then

∆β ∆γ ∆δ
∆HT° = ∆Ho + ∆αT + 2 T2 + 3 T3 – T … (3.77)

So with Equation (3.76) or (3.77), we can estimate the standard heat of reaction at any
temperature T. In the above equations, T is the absolute temperature.
Adiabatic flame temperature :
A process in which heat can leave or enter a system is called an adiabatic process. Here we
consider adiabatic process as the one in which the heat is contained within a system without loss
or gain of heat to or from the outside source (surroundings).
A reaction is said to an adiabatic reaction if there is no heat interaction between the system
and the surroundings.
If the adiabatic reaction is exothermic, the heat released during the reaction will be utilised to
increase the enthalpy of the products since no heat is removed to keep the temperature constant
so the temperature of the product system rises. The temperature attained by the product steam is
called the adiabatic temperature of the reaction. In this case, the total heat content/enthalpy of the
product steam is equal to the heat content/enthalpy of reactants plus the heat of reaction.
When a fuel is burned in air or oxygen under adiabatic conditions, the temperature attained
by the product stream is called the adiabatic flame temperature. The maximum adiabatic flame
temperature is attained when the fuel is burned with theoretically/ stoichiometrically required
amount of pure oxygen. When the fuel is burned with air or excess oxygen then the adiabatic
flame temperature attained is less than that attained with pure oxygen. Even in well-insulated
furnaces, the heat is lost due to radiation heat transfer from the surface and hence, the actual
flame temperature is less than the adiabatic flame temperature.
SOLVED EXAMPLES

Example 3.1 :
Determine the molar volume of methane at 600 bar and 27oC (300 K) by using :
(i) the ideal gas equation,
(ii) the van der Waals equation.
The van der Waals constants a and b are 0.2285 N·m4/mol2 and 4.27 × 10–5 m3/mol,
respectively.
(iii) the Redlich-Kwong equation. The critical temperature and pressure are 191.1 K and
46.4 bar, respectively.
Chemical Engineering Thermodynamics - I 3.36 P-V-T Behaviour and Heat Effects

Solution :
(i) To find the molar volume using the ideal gas equation :
RT
V = P ... for 1 mol of gas

where R = 8.31451 m3·Pa/(mol·K)


T = 300 K and P = 600 bar = 600 × 105 Pa
8.31451 × 300
∴ V = = 4.16 × 10–5 m3/mol ... Ans. (i)
600 × 105
(ii) To find the molar volume using the van der Waals equation of state :
P + a  (V – b) = RT ... for 1 mol of gas
 V 2
N
where a = 0.2285 N·m4/mol2 = 0.2285 m2 · m6/mol2 = 0.2285 Pa·m6/mol2

b = 4.27 × 10–5 m3/mol

∴ 600 × 105 + 0.2285 (V – 4.27 × 10–5) = 8.31451 × 300 = 2494.35


 V2 
Volume can be determined by adopting a trial and error procedure.
R.H.S. = 2494.35

L.H.S. = 600 × 105 + V2  (V – 4.27 × 10–5)


0.2285
 
Assume a value of V and evaluate the R.H.S. of the van der Waals equation. See whether
R.H.S. ≈ L.H.S. or not. If not, then repeat the procedure till we get R.H.S. ≈ L.H.S.
The ideal gas volume may be chosen for the first trial.
Trial 1 : Put V = 4.16 × 10–5 m3/mol

L.H.S. = 600 × 105 +


0.2285 
(4.16 × 10–5 – 4.27 × 10–5) = – 211.4
 (4.16 × 10–5)2
∴ L.H.S. ≠ R.H.S.
Trial 2 : Put V = 5 × 10–5 m3/mol : L.H.S. = 1105.22, R.H.S. = 2494.35
Trial 3 : Put V = 6.5 × 10–5 m3/mol : L.H.S. = 2505, R.H.S. = 2494.35
Trial 4 : Put V = 6.45 × 10–5 m3/mol : L.H.S. = 2505, R.H.S. = 2494.35
Trial 5 : Put V = 6.44 × 10–5 m3/mol : L.H.S. = 2497.57, R.H.S. = 2494.35
∴ For V = 6.44 × 10–5 m3/mol, L.H.S. (2497.57) ≈ R.H.S. (2494.35)

Molar volume of methane gas using = 6.44 × 10–5 m3/mol ... Ans. (ii)
 the van der Waals equation of state 
Chemical Engineering Thermodynamics - I 3.37 P-V-T Behaviour and Heat Effects

OR
Adopt a procedure of iterative calculations.
The iterative calculations of methane gas volume can be initiated with V = ideal gas volume.
For this purpose, the van der Waals equation can be rearranged as
RT
V = +b
P + a 
 V 2
RT
Vcalculated = +b
P + a 
 V2 
 assumed

8.31451 × 300
0.2285 + 4.27 × 10
Vcalculated = –5

600 × 10 + 2
5
Vassumed
Iteration 1 : Vassumed = Volume predicted by the ideal gas law
= 4.16 × 10–5 m3/mol
8.31451 × 300
Vcalculated = 0.2285 + 4.27 × 10–5
600 × 10 +
5
(4.16 × 10–5)2
= 5.57 × 10–5 m3/mol
Vcalculated (5.57 × 10–5) ≠ Vassumed (4.16 × 10–5)
∴ Now assume V = Vcalculated = 5.57 × 10–5 m3/mol
Iteration 2 : Vassumed = 5.57 × 10–5 m3/mol
8.31451 × 300
Vcalculated = 0.2285 + 4.27 × 10–5
600 × 10 +
5
(5.57 × 10–5)
= 6.13 × 10–5 m3/mol
Vassumed (5.57 × 10–5) ≠ Vcalculated (6.13 × 10–5)
Iteration 3 : Put V = 6.13 × 10–5 m3/mol for Vassumed on the R.H.S. of the rearranged
equation
8.31451 × 300
V = 0.2289 + 4.27 × 10–5
600 × 10 +
5
(6.13 × 10–5)
= 6.33 × 10–5 m3/mol
Chemical Engineering Thermodynamics - I 3.38 P-V-T Behaviour and Heat Effects

Iteration 4 : Put V = 6.33 × 10–5 m3/mol for Vassumed on the R.H.S. of the rearranged
equation
8.31451 × 300
V = 0.2289 + 4.27 × 10–5
600 × 10 +
5
(6.33 × 10–5)2
= 6.4 × 10–5 m3/mol
Iteration 5 : Put V = 6.4 × 10–5 m3/mol
8.31451 × 300
V = 0.2289 + 4.27 × 10–5
600 × 10 +
5
(6.4 × 10–5)2
= 6.422 × 10–5 m3/mol
Iteration 6 : Put V = 6.422 × 10–5 m3/mol
8.31451 × 300
V = 0.2289 + 4.27 × 10–5
600 × 10 +
5
(6.422 × 10–5)2
= 6.43 × 10–5 m3/mol
∴ Vcalculated (6.43 × 10–5 m3/mol) ≈ Vassumed (6.422 × 10–5 m3/mol)
Hence, calculations can be terminated at this stage.
Therefore, molar volume of methane gas = 6.43 × 10–5 m3/mol ... Ans. (ii)
(iii) To calculate the molar volume using the Redlich-Kwong equation :
First we will evaluate the constants a and b.
Given : Pc = 46.4 bar = 46.4 × 105 Pa and Tc = 191.1 K
2.5
0.4278 R2 Tc 0.4278 (8.31451)2 (191.1)2.5
a = Pc =
46.4 × 105

= 3.22 m6·Pa/mol2 (= 3.22 N·m4/mol2)


0.0867 RTc 0.0867 × 8.31451 × 191.1
b = Pc =
46.4 × 105

= 2.97 × 10–5 m3/mol


The RK equation is
RT a
P = V – b – T0.5 V (V + b)

where P = 600 bar = 600 × 105 Pa


T = 300 K
Chemical Engineering Thermodynamics - I 3.39 P-V-T Behaviour and Heat Effects

8.31451 × 300 3.22


∴ 600 × 105 = –
V – 2.97 × 10 –5 (300) V (V + 2.97 × 10–5)
0.5

2494.35 0.186
600 × 105 = –
V – 2.97 × 10 –5 V (V + 2.97 × 10–5)
L.H.S. = 600 × 105 = 6 × 107
2494.35 0.186
R.H.S. = –
V – 2.97 × 10 –5 V (V + 2.97 × 10–5)
We will find V by a trial and error procedure.
To find V, use the ideal gas volume (i.e., volume predicted by the ideal gas law) as an initial
gauss value and evaluate the R.H.S.
Check for L.H.S. = R.H.S. If not, assume a new value of V and repeat the procedure till
we get L.H.S. ≈ R.H.S.
Trial 1 : Take V = 4.16 × 10–5 m3/mol ... assumed value for V
2494.35 0.186
R.H.S. = –
4.16 × 10 – 2.97 × 10
–5 –5 4.16 × 10 (4.16 × 10–5 + 2.97 × 10–5)
–5

= 1.47 × 108
∴ L.H.S. (6 × 107) ≠ R.H.S. (1.47 × 108)
Trial 2 : For V = 5 × 10–5 m3/mol : R.H.S. = 7.64 × 107, L.H.S. = 6 × 107
Trial 3 : For V = 5.5 × 10–5 m3/mol : R.H.S. = 5.87 × 107, L.H.S. = 6 × 107
Trial 4 : For V = 5.4 × 10–5 m3/mol : R.H.S. = 6.15 × 107, L.H.S. = 6 × 107
Trial 5 : For V = 5.45 × 10–5 m3/mol : R.H.S. = 6.005 × 107, L.H.S. = 6 × 107
∴ For V = 5.45 × 10–5 m3/mol : L.H.S. ≈ R.H.S.
Vcalculated in the fifth iteration is the final root. Therefore,
Molar volume of methane gas using the Redlich-Kwong equation = 5.45 × 10–5 m3/mol
... Ans. (iii)
Example 3.2 :
A vessel of volume 0.03 m3 contains 0.5 kg gaseous ammonia at a constant temperature of
338 K. The critical pressure and temperature are 112.8 bar and 405.5 K, respectively. Calculate
the pressure developed by this gas using the Redlich-Kwong equation.
Solution :
First evaluate the constants a and b.
Tc = 405.5 K and Pc = 112.8 bar = 112.8 × 105 N/m2
2.5
0.4278 R2 Tc 0.4278 × (8.31451)2 × (405.5)2.5
a = Pc =
112.8 × 105
= 8.68 N·m4·K0.5/mol2
Chemical Engineering Thermodynamics - I 3.40 P-V-T Behaviour and Heat Effects

= 8.68 × 10–3 kN·m4·K0.5/mol2

= 8.68 × 10–3  m2  · m6·K0.5 /(10–3 kmol)2


kN
 
= 8680 kPa·m6·K0.5/(kmol)2
0.0867 RTc 0.0867 × 8.31451 × 405.5
b = Pc =
112.8 × 105
= 2.59 × 10–5 m3/mol = 2.59 × 10–5 m3/(10–3 kmol)
= 2.59 × 10–2 m3/kmol
The Redlich-Kwong equation is
RT a
P = V – b – T0.5 ·V (V + b)

where V is in m3/kmol
Amount of NH3 gas = 0.5 kg, Mol. wt. of NH3 = 17
0.5
∴ Moles of NH3 = 17 = 0.0294 kmol

Vessel volume = 0.03 m3


0.03
∴ V (molar volume) = 0.0294 = 1.02 m3/kmol

T = 338 K
8.31451 × 338 8680
P = –
(1.02 – 2.59 × 10 ) (338) × 1.02 (1.02 + 2.59 × 10–2)
–2 0.5

= 2384 kPa = 23.84 bar ... Ans.


Example 3.3 :
Calculate the pressure developed by 1 kmol of ammonia gas contained in a vessel of 0.6 m3
volume at a constant temperature of 200°C (473 K) by using
(i) the ideal gas equation,
(ii) the van der Waals equation. The van der Waals constants a and b are 0.4233 N·m4/mol2
and 3.73 × 10–5 m3/mol, respectively.
(iii) the Redlich-Kwong equation. Pc = 112.8 bar and Tc = 405.5 K.
Solution :
(i) P using the ideal gas equation :
nRT
P = V
where n = 1 kmol, R = 8.31451 kPa·m3/(kmol·K), T = 473 K and V = 0.6 m3
1 × 8.31451 × 473
∴ P = 0.60 = 6554.6 kPa
= 65.546 bar ≈ 65.55 bar ... Ans. (i)
Chemical Engineering Thermodynamics - I 3.41 P-V-T Behaviour and Heat Effects

(ii) P using the van der Waals equation :


P + a  (V – b) = RT
 V 2
RT a
∴ P = V – b – V2

where a = 0.4233 N·m4/mol2 = 0.4233 × 10–3 kN·m4/mol2

= 0.4233 × 10–3  m2  · m6 (10–3 kmol)2


kN
  /
= 423.3 kPa·m6/(kmol)2
b = 3.73 × 10–5 m3/mol = 3.73 × 10–2 m3/kmol
Moles of NH3 gas = 1 kmol
Volume of vessel = 0.6 m3
V in the van der Waals equation is in m3/kmol.
0.60
∴ V = 1 = 0.60 m3/kmol
Substituting the numerical values yields
8.31451 × 473 423.3
P = –
(0.6 – 3.73 × 10–2) (0.6)2
= 58.13 kPa = 58.13 bar ... Ans. (ii)
(iii) P using the Redlich-Kwong equation :
2.5
0.4278 R2 Tc
a = Pc
where Tc = 405.5 K and Pc = 112.8 bar = 112.8 × 105 N/m2
0.4278 × (8.31451)2 (405.5)2.5
∴ a =
112.8 × 105
= 8.68 N·m4K0.5/mol2 = 8680 kPa·m6·K0.5/kmol2
0.0867 RTc
b = Pc
0.0867 × 8.31451 × 405.5
=
112.8 × 105
= 2.59 × 10–5 m3/mol = 2.59 × 10–2 m3/kmol
The Redlich-Kwong equation is
RT a
P = V – b – T0.5 V (V + b)
where V = 0.6 m3/1 kmol = 0.6 m3/kmol
8.31451 × 473 8680
∴ P = –
(0.6 – 2.59 × 10 ) (473) × 0.60 (0.6 + 2.59 × 10–2)
–2 0.5

= 5787.6 kPa = 57.876 bar ≈ 57.88 bar ... Ans. (iii)


Chemical Engineering Thermodynamics - I 3.42 P-V-T Behaviour and Heat Effects

Example 3.4 :
Calculate the molar volume of saturated vapour of methyl chloride at 333 K. Assume that
methyl chloride follows the Redlich-Kwong equation of state.
The saturation pressure of methyl chloride at 333 K is 13.76 bar. The critical constants of
methyl chloride are Tc = 416.3 K and Pc = 66.8 bar.
Solution :
The critical constants of the RK equation of state are given by
0.5
0.4278 R2 Tc
a = Pc m6·Pa·K0.5/mol2 or N·m4·K0.5/mol2
0.0867 RTc
and b = Pc m3/mol

R = 8.31451 m3·Pa/(mol·K), Tc = 416.3 K and Pc = 66.8 bar = 66.8 × 105 Pa


0.4278 (8.31451)2 (416.3)2.5
a = = 15.655 m6·Pa·K0.5/mol2
66.8 × 105
0.0867 (8.31451) (416.3)
b = = 4.49 × 10–5 m3/mol
66.8 × 105
The ideal gas volume is
RT
V = P
where P = 13.76 bar = 13.76 × 105 Pa and T = 333 K
8.31451 × 333
V =
13.76 × 105
= 2.012 × 10–3 m3/mol
The Redlich-Kwong equation is
RT a
P = V – b – T0.5 V (V + b)
Calculation of V for given P and T requires a trial and error method. For an iterative
calculation, the above equation can be rearranged as
RT a (V – b)
V = P + b – T0.5 PV (V + b)
The iterative calculation of saturated vapour volume can be initiated by taking (assuming) V
on the R.H.S. of the above equation equal to V obtained by using the ideal gas law. The
calculations are terminated when the assumed and calculated V values are approximately equal.
[If Vassumed ≠ Vcalculated in the first iteration, then the Vcalculated in the first iteration is to be used
as Vassumed in the second iteration.]
8.31451 × 333 15.055 (V – 4.49 × 10–5)
Iteration 1 : V = + 4.49 × 10 –5 –
13.76 × 105 (333)0.5 × 13.76 × 10–5 V (V + 4.49 × 10–5)
6.2346 × 10–7 (V – 4.49 × 10–5)
V = 2.0569 × 10–5 – ... (1)
V (V + 4.49 × 10–5)
Chemical Engineering Thermodynamics - I 3.43 P-V-T Behaviour and Heat Effects

Take V = 2.012 × 10–3 m3/mol (on the R.H.S. of the above equation)
6.2346 × 10–7 (2.012 × 10–3 – 4.49 × 10–5)
V = 2.0569 × 10–3 –
2.012 × 10–3 (2.012 × 10–3 + 4.49 × 10–5)
V = 1.761 × 10–3 m3/mol
We see that the Vassumed is not equal to Vcalculated.
Iteration 2 : Take V = 1.761 × 10–3 m3/mol. Using Equation (1),
6.2346 × 10–7 (1.761 × 10–3 – 4.49 × 10–5)
V = 2.0569 × 10–3 –
1.761 × 10–3 (1.761 × 10–3 + 4.49 × 10–5)
= 1.72 × 10–3 m3/mol
Vassumed (1.761 × 10–3) ≠ Vcalculated (1.72 × 10–3)
Iteration 3 : Take V = 1.72 × 10–3 m3/mol. Using Equation (1),
6.2346 × 10–7 (1.72 × 10–3 – 4.49 × 10–5)
V = 2.0569 × 10–3 –
1.72 × 10–3 (1.72 × 10–3 + 4.49 × 10–5)
= 1.713 × 10–3 m3/mol
Iteration 4 : Take V = 1.713 m3/mol. Using Equation (1),
6.2346 × 10–7 (1.713 × 10–3 – 4.49 × 10–5)
V = 2.0569 × 10–3 –
1.713 × 10–3 (1.713 × 10–3 + 4.49 × 10–5)
= 1.711 × 10–3 m3/mol
We see that the Vassumed (1.713 × 10–3 m3/mol) ≈ Vcalculated (1.711 × 10–3 m3/mol). Hence,
the calculations are terminated at this stage. Vcalculated is the final root. Therefore,
Molar volume of saturated vapour of methyl chloride = 1.711 × 10–3 m3/mol ... Ans.
Extra : To calculate the molar volume of saturated liquid :
Here assume V > b (= 4.49 × 10–5)
6.2346 × 10–7 (V – 4.49 × 10–5)
V = 2.0569 × 10–3 –
V (V + 4.49 × 10–5)
For V = 7.133 × 10–5 ... Vcalculated = 6.94 × 10–5 ... difference = 1.93 × 10–6
For V = 7.128 × 10–5 ... Vcalculated = 7.09 × 10–5 ... difference = 3.8 × 10–7
For V = 7.125 × 10–3 ... Vcalculated = 7.179 × 10–5 ... difference = 5.4 × 10–7
For V = 7.129 × 10–5 ... Vcalculated = 7.06 × 10–5 ... difference = 6.9 × 10–7
For V = 7.127 × 10–5 ... Vcalculated = 7.12 × 10–5 ... difference = 7 × 10–8
Since Vassumed ≈ Vcalculated,
Vcalculated = 7.12 × 10–5 is the final root
Molar volume of saturated methyl chloride liquid = 7.12 × 10–5 m3/mol ... Ans.
Chemical Engineering Thermodynamics - I 3.44 P-V-T Behaviour and Heat Effects

Example 3.5 :
Calculate the molar volumes of saturated liquid and saturated vapour of n-octane at
427.85 K. The saturation pressure of n-octane at 427.85 K is 0.215 MPa. Assume that n-octane
follows the van der Waals equation of state.
Data : The van der Waals constants a and b are 3.789 Pa·m6/mol2 and 2.37 × 10–4 m3/mol,
respectively.
Solution :
T = 427.85 K, P = 0.215 MPa = 0.215 × 106 Pa,
a = 3.789 Pa·m6/mol2 and b = 2.37 × 10–4 m3/mol
(i) To find the molar volume of saturated n-octane vapour : First calculate the molar
volume using the ideal gas equation.
RT
V = P ... for 1 mol of gas
8.31451 × 427.85
= = 0.01655 m3/mol
0.215 × 106
The van der Waals equation is
P + a  (V – b) = RT ... for 1 mol
 V 2
Calculation of V from given P and T requires a trial and error procedure. For an iterative
calculation, the above equation can be rearranged as
RT
V = +b
P + a 
 V 2
The iterative calculations of vapour volume can be initiated by assuming V (on the R.H.S. of
the above equation) equal to the volume predicted by the ideal gas law. Therefore, use Vobtained
using the ideal gas equation (ideal gas volume) for V on the R.H.S. for the first iteration, and
calculate V (the L.H.S.). If Vcalculated is not (approximately) equal to Vassumed, then use
Vcalculated for V for the second iteration and calculate V. This iterative calculation procedure is
continued till the difference between the assumed and calculated V values is very small.
Here is the procedure :
RT kPa·m3 Pa·m3
V = + b , R = 8.31451 = 8.31451
P + a  kmol·K mol·K
 V 2

8.31451 × 427.85
V = + 2.37 × 10–4
0.215 × 106 + 3.789
 V2 
3557.36
V = + 2.37 × 10–4
0.215 × 106 + 3.789
 V2 
Chemical Engineering Thermodynamics - I 3.45 P-V-T Behaviour and Heat Effects

First iteration : V (on the R.H.S.) = V (ideal gas) = 0.01655 m3/mol


3557.36
V = + 2.37 × 10–4
0.215 × 106 + 3.789 
 (0.01655)2
= 0.01578 m3/mol ... Vcalculated
Vassumed ≠ Vcalculated
Second iteration : V (on the R.H.S.) = 0.01578 ... Vcalculated in the first iteration
3557.36
V = + 2.37 × 10–4
0.215 × 106 + 3.789 
 (0.01578)2
= 0.01569 m3/mol
Third iteration : Take V = 0.01569 m3/mol
3557.36
V = + 2.37 × 10–4
0.215 × 106 + 3.789 
 (0.01569)2
= 0.01568 m3/mol
Vassumed = 0.01569 m3/mol and Vcalculated = 0.01568 m3/mol
We see that by the third trial, the difference between the assumed and calculated V values is
very small (1 × 10–5). Hence, calculations can be terminated at this stage. Vcalculated is the final
root. Therefore,
Molar volume of saturated = 0.01568 m3/mol = 15.68 × 10–3 m3/mol ... Ans.
 n-octane vapour 
To find the molar volume of saturated liquid : The calculations can be initiated with
assuming V on the R.H.S. of the equation equal to b.
3557.36
V = 3.789 + b
0.215 × 106 + V2

First iteration : V on the R.H.S. = b = 2.37 × 10–4 m3/mol


3557.36
V = 3.789 + 2.37 × 10–4
0.215 × 10 +
6
(2.37 × 10–4)2
= 2.8956 × 10–4 ≈ 2.90 × 10–4
Vassumed ≠ Vcalculated
Second iteration : Take V = 2.90 × 10–4 m3/mol
3557.36
V = 3.789 + 2.37 × 10–4 = 3.16 × 10–4 m3/mol
0.215 × 10 +
6
(2.90 × 10–4)2
Chemical Engineering Thermodynamics - I 3.46 P-V-T Behaviour and Heat Effects

Third iteration : Take V = 3.16 × 10–4 m3/mol ... Vcalculated = 3.30 × 10–4 m3/mol
Fourth iteration : Take V = 3.30 × 10–4 m3/mol ... Vcalculated = 3.39 × 10–4 m3/mol
Fifth iteration : Take V = 3.39 × 10–4 m3/mol ... Vcalculated = 3.44 × 10–4 m3/mol
Sixth iteration : Take V = 3.44 × 10–4 m3/mol ... Vcalculated = 3.49 × 10–4 m3/mol
Seventh iteration : Take V = 3.49 × 10–4 m3/mol ... Vcalculated = 3.506 × 10–4 m3/mol
∴ Vassumed (3.49 × 10–4 m3/mol) ≈ Vcalculated (3.506 × 10–4 m3/mol)
We see that the change in the volume between the sixth and seventh iterations (i.e., the
change in the assumed and calculated V values in the seventh iteration) is very small (1.6 × 10–6).
Hence, the calculations can be terminated at this stage. Vcalculated is the final root. Therefore,
Molar volume of saturated n-octane liquid = 3.506 × 10–4 m3/mol ... Ans.
Example 3.6 :
Calculate the molar volume and compressibility factor of isopropyl alcohol vapour at 473 K
and 10 bar. Assume that isopropyl alcohol follows the virial equation of state. The virial
coefficients B and C are – 3.88 × 10–4 m3/mol and – 2.6 × 10–8 m6/mol2, respectively.
Solution :
The virial equation truncated to three terms is
PV B C
Z = RT = 1 + V + V2
This equation can be rearranged as
V = P 1 + V + V2
RT B C
 
where R = 8.31451 m3·Pa/(mol·K), T = 473 K, P = 10 bar = 10 × 105 Pa,
B = – 3.88 × 10–4 m3/mol and C = – 2.6 × 10–8 m6/mol2
8.31451 × 473  3.88 × 10–4 2.6 × 10–8
V = 1– –
10 × 105  V V2 

V = 3.933 × 10–3 1 –
3.88 × 10–4 2.6 × 10–8
– ... (A)
 V V2 
V is determined by a trial and error procedure.
To initiate the iterative calculation, use the ideal gas volume as an initial guess value.
RT 8.31451 × 473
V = P = = 3.933 × 10–3 m3/mol
10 × 105
Iteration 1 : Put V = 3.933 × 10–3 m3/mol on the R.H.S. of Equation (A) and evaluate V.
 3.88 × 10–4 2.6 × 10–8 
V = 3.933 × 10–3 1 – – 
 3.933 × 10–3 (3.933 × 10–3)2
= 3.54 × 10–3 m3/mol
We see that Vassumed (3.933 × 10–3) ≠ Vcalculated (3.54 × 10–3).
Hence, continue the calculations with a new value of V.
Use V = 3.54 × 10–3 m3/mol for V on the R.H.S. of Equation (A) in the second iteration.
Chemical Engineering Thermodynamics - I 3.47 P-V-T Behaviour and Heat Effects

Iteration 2 : Put V = 3.54 × 10–3 m3/mol (Vcalculated in the first iteration) on the R.H.S. of
Equation (A) and evaluate V on the L.H.S.
 3.88 × 10–4 2.6 × 10–8 
V = 3.933 × 10–3 1 – – 
 3.54 × 10 –3 (3.54 × 10–3)2
= 3.49 × 10–3 m3/mol
Vassumed ≠ Vcalculated
Use V = 3.49 × 10–3 m3/mol for V on the R.H.S. of Equation (A) in the third iteration.
Iteration 3 : Put V = 3.49 × 10–3 m3/mol

V = 3.933 × 10–3 1 –
 3.88 × 10–4 2.6 × 10–8 
(3.49 × 10–3)2

 3.49 × 10 –3

= 3.487 × 10–3 m3/mol


As Vassumed (3.49 × 10–3) ≈ Vcalculated (3.487 × 10–3), the calculations can be terminated at
this stage. Vcalculated is the final root. Therefore,
Molar volume of isopropyl alcohol vapour = 3.487 × 10–3 m3/mol ... Ans.
The compressibility factor is calculated as
PV V Actual volume of gas
Z = RT = (RT/P) = Volume predicted by ideal gas law

10 × 105 × 3.487 × 10–3 3.487 × 10–3


= Or = 0.8866 ≈ 0.887 ... Ans.
8.31451 × 473 3.933 × 10–3
Example 3.7 :
Calculate the molar volume of saturated vapour of n-octane at 427.85 K and 0.215 MPa.
Assume that n-octane follows the Soave-Redlich-Kwong equation of state.
Take Tc = 569.4 K, Pc = 24.97 bar and ω = 0.398 for n-octane.
Solution :
The Soave-Redlich-Kwong equation is
RT aα
P = V – b – V (V + b)
2
0.4278 R2 Tc 0.4278 (8.31451)2 (569.4)2
where a = Pc = = 3.84 m6·Pa/mol2
24.97 × 105
0.0867 RTc 0.0867 × 8.31451 × 569.4
b = Pc = = 1.644 × 10–4 m3/mol
24.97 × 105
ω = 0.398
S = 0.48 + 1.574 ω – 0.176 ω2
= 0.48 + 1.574 (0.398) – 0.176 (0.398)2 = 1.078
Chemical Engineering Thermodynamics - I 3.48 P-V-T Behaviour and Heat Effects

T 427.85
α = 1 + S (1 – T r) , Tr = T = 569.4 = 0.7514
c

α = 1 + 1.078 (1 – 0.7514)
∴ α = 1.308
To calculate V, the SRK equation can be rearranged as
RT aα (V – b)
V = P + b – PV (V + b)

where T = 427.85 K, R = 8.31451 m3·Pa/(mol·K), P = 0.215 MPa = 0.215 × 106 Pa


Find V using the ideal gas equation :
RT
V = P = 8.31451 × 427.85/0.215 × 106

= 0.01655 m3/mol
= 16.55 × 10–3 m3/mol
Substituting the numerical values in the rearranged SRK equation, we get
8.31451 × 427.85 3.84 × 1.308 (V – 1.644 × 10–4)
V = + 1.644 × 10–4 –
0.215 × 10 6 0.215 × 106 V (V + 1.644 × 10–4)
2.336 × 10–5 (V – 1.644 × 10–4)
V = 0.016714 – ... (A)
V (V + 1.644 × 10–4)
Iteration calculations of V can be started with V on the R.H.S. of Equation (A) = V predicted
by the ideal gas law.
Iteration 1 : Assume V = 16.55 × 10–3 m3/mol
2.33 × 10–5 (16.55 × 10-3 – 1.644 × 10–4)
V = 0.016714 –
16.55 × 10–3 (16.55 × 10–3 + 1.644 × 10–4)
= 15.33 × 10–3 m3/mol
Since Vcalculated ≠ Vassumed, go for a second iteration.
Iteration 2 : Assume V = 15.33 × 10–3 m3/mol
2.336 × 10–5 (15.33 × 10–3 – 1.644 × 10–4)
V = 0.016714 –
15.33 × 10–3 (15.33 × 10–3 + 1.644 × 10–4)
= 15.22 × 10–3 m3/mol
Vcalculated ≠ Vassumed
Iteration 3 : Assume V = 15.22 m3/mol
Vcalculated = 15.21 × 10–3 m3/mol
As Vassumed (15.22 m3/mol) is almost equal to Vcalculated (15.21 × 10–3 m3/mol), the
calculations are terminated at this stage. Vcalculated is the final root. Therefore,
Molar volume of saturated n-octane vapour = 15.22 × 10–3 m3/mol ... Ans.
Chemical Engineering Thermodynamics - I 3.49 P-V-T Behaviour and Heat Effects

Example 3.8 :
It is required to store 100 kg methane in a tank at 298 K and 2 MPa. Calculate the capacity
of the tank using the Soave-Redlich-Kwong equation of state.
For methane : Tc = 190.7 K, Pc = 46.41 bar and ω = 0.011
Solution :
Find the ideal gas volume.
RT
V = P

where R = 8.31451 m3·Pa/(mol·K), P = 2 MPa = 2 × 106 Pa and T = 298 K


8.31451 × 298
∴ V = = 1.24 × 10–3 m3/mol
2 × 106
Find the constants a and b of the RK equation.
2
0.4278 R2 Tc 0.4278 (8.31451)2 (190.7)2
a = Pc = = 0.2317 m6·Pa/mol2
46.41 × 105
0.0867 RTc 0.0867 × 190.7
b = Pc = = 3.56 × 10–6 m3/mol
46.41 × 105
S = 0.48 + 1.574 ω – 0.176 ω2
= 0.48 + 1.574 × 0.011 – 0.176 (0.011)2 = 0.4973
T 298
α = 1 + S ( 1 – T r) , Tr = T = 190.7 = 1.563
c

= 1 + 0.4973 (1 – 1.563)
∴ α = 0.87567 ≈ 0.876
The SRK equation can be rearranged as given below for an iterative calculation of V.
RT aα (V – b)
V = p + b – PV (V + b)
8.31451 × 298 0.2317 × 0.876 (V – 3.56 × 10–6)
= + 3.56 × 10 –6 –
2 × 106 2 × 106 × V (V + 3.56 × 10–6)
1.015 × 10–7 (V – 3.56 × 10–6)
= 1.2436 × 10–3 – ... (A)
V (V + 3.56 × 10–6)
Iteration 1 : Assume the ideal gas volume (1.24 × 10–3 m3/mol) for V on the R.H.S. of
Equation (A) and evaluate V on the LHS.
V = 1.24 × 10–3 m3/mol
1.015 × 10–7 (1.24 × 10–3 – 3.56 × 10–6)
V = 1.2436 × 10–3 –
1.24 × 10–3 (1.24 × 10–3 + 3.56 × 10–6)
= 1.16 × 10–3 m3/mol
∴ Vassumed (1.24 × 10–3) ≠ Vcalculated (1.16 × 10–3)
Chemical Engineering Thermodynamics - I 3.50 P-V-T Behaviour and Heat Effects

Iteration 2 : Assume
V on the RHS = 1.16 × 10–3 m3/mol
1.015 × 10–7 (1.16 × 10–3 – 3.56 × 10–6)
V = 1.2436 × 10–3 –
1.16 × 10–3 (1.16 × 10–3 + 3.56 × 10–6)
= 1.1566 × 10–3 ≈ 1.157 × 10–3 m3/mol
Iteration 3 : V = 1.157 × 10–3 m3/mol
1.015 × 10–7 (1.157 × 10–3 – 3.56 × 10–6)
V = 1.2436 × 10–3 –
1.157 × 10–3 (1.157 × 10–3 + 3.56 × 10–6)
= 1.156 × 10–3 m3/mol
We see that Vassumed (1.157 × 10–3) ≈ Vcalculated (1.156 × 10–3). Hence, calculations can be
terminated at this stage. Vcalculated in the third iteration is the final root. Therefore,
Molar volume of methane = 1.156 × 10–3 m3/mol
We will make use of this molar volume to calculate the capacity of the tank.
Amount of methane = 100 kg
Mol. wt. of methane = 16
100
∴ Moles of methane = 16 = 6.25 kmol = 6250 mol
Molar volume of methane = 1.156 × 10–3 m3/mol
Volume of the tank
= Moles of methane in the tank
∴ Volume of the tank (capacity) = 1.156 × 10–3 × Moles of methane in the tank
m3
= 1.156 × 10–3 mol × 6250 mol

= 7.225 m3
∴ The capacity of the tank is 7.225 m3 ... Ans.
Example 3.9 :
Calculate the standard heat of the following reaction at 298 K.
C5H12 (g) + 8 O2 (g) –→ 5 CO2 (g) + 6 H2O (l)
The standard heats of formation of the components are :
o
Component ∆Hf 298 kJ/mol
CO2 (g) – 393.509
H2O (g) – 241.818
C5H12 (g) – 146.76
The latent heat of vaporisation of water at 298 K is 43.967 kJ/mol.
Solution :
C5H12 (g) + 8 O2 (g) –→ 5 CO2 (g) + 6 H2O (l)
Chemical Engineering Thermodynamics - I 3.51 P-V-T Behaviour and Heat Effects

o
To find the standard heat of reaction of the above reaction, we are provided with ∆Hf 298 for
o o
C5H12 (g) and H2O (g). But we need ∆Hf 298 for H2O (l). To calculate ∆Hf 298 for H2O (l), we are
provided with the enthalpy of vaporisation of water at 298 K.
Latent heat of vaporisation = Enthalpy of vaporisation
 of water at 298 K   of water at 298 K 
= 43.967 kJ.mol
H2O (g) –→ H2O (l) ... Phase change
For the above phase change,
o
∆H298 = –ve of the enthalpy of vaporisation at 298 K
= – 43.967 kJ/mol
o o o
∆H298 = ∆Hf 298 [H2O (l)] – ∆Hf 298 [H2O (g)]
o o o
∴ ∆Hf 298 [H2O (l)] = ∆H298 + ∆Hf 298 [H2O (g)]
= – 43.967 + (– 241.818)
= – 285.785 kJ/mol
Now, we know the standard heat of formation of all the components participated in the
reaction.
o
∆H298 = Standard heat of reaction at 298 K
= Σ [ni ∆Hof 298 (i) ]products – Σ [ni ∆Hof 298 (i) ]reactants
where ni is the stoichiometric coefficient of ith component.
o o o
∆H298 = {5 × ∆H f 298 [CO2] + 6 ∆Hf 298 [H2O (l)] }
o o
– {1 × ∆H f 298 [C5H12] + 8 × ∆Hf 298 [O2] }
= 5 × (– 393.509) + 6 (– 285.785) – 1 × (– 146.76) – 8 × (0)
= – 3535.495
≈ – 3535.5 kJ/mol C5H12 or – 3535.5 kJ ... Ans.
Example 3.10 :
A gas mixture containing 20% CO and 80% N2 by volume is burned with 100%
excess air. Initially, both air and gas are at 298 K. The standard heat of reaction at 298 K is
– 283.178 kJ/mol CO. The heat capacities of the components are represented by
o
Cp = a + bT + cT2
o
where Cp is in J/(mol·K) and the constants a, b and c are :
Component a b c
CO2 26.54 42.45 × 10–3 – 14.298 × 10–6
O2 25.61 13.26 × 10–3 – 4.208 × 10–6
N2 27.03 5.815 × 10–3 – 0.289 × 10–6
Calculate the theoretical flame temperature attained.
Chemical Engineering Thermodynamics - I 3.52 P-V-T Behaviour and Heat Effects

Solution :
Basis : 1 mol of gas mixture.
It contains 0.20 mol CO and 0.8 mol N2.
1
CO + 2 O2 –→ CO2

Amount of CO reacted = CO fed = 0.20 mol


1/2
O2 reacted = 1 × 0.20 = 0.10 mol

1/2
O2 theoretically required for 0.20 mol CO fed = 1 × 0.20 = 0.10 mol

Excess air supplied = 100%


Excess O2 = Excess air = 100%

O2 in air supplied = O2 theoretical 1 +


% excess
 100 

= 0.10 1 + 100 = 0.20 mol


100
 
O2 unreacted = 0.20 – 0.10 = 0.10 mol
79
N2 in the air supplied = 21 × 0.20 = 0.7524 mol

N2 in the reactant stream = 0.7524 + 0.80 = 1.5524 mol


Reactant stream : 0.2 mol CO + 1.5524 mol N2 + 0.20 mol O2
Reactant stream is at T = 298 K (both air and gas mixture are at T = 298 K)
∆H1 = Enthalpy of reactant stream at 298 K over 298 K
0 ... as both temperatures are the same.
o
∆H298 = – 283.178 kJ/mol CO reacted
= – 283178 J/mol CO reacted
Amount of CO reacted = 0.20 mol
Total standard heat of reaction at 298 K is
o 283178
Total ∆H298 = – 1 × 0.20 = – 56635.6 J

N2 in the product stream = N2 in the reactant stream … it being inert


= 1.5524 mol
Product stream : 0.10 mol O2 + 1.5524 mol N2 + 0.20 mol CO2
Let the theoretical flame temperature be T K. The product stream leaves the combustion
chamber at T K.
Chemical Engineering Thermodynamics - I 3.53 P-V-T Behaviour and Heat Effects

o
Cp for CO2 = 26.54 + 42.45 × 10–3 T – 14.298 × 10–6 T2 ... J/(mol·K)
o
Cp for O2 = 25.61 + 13.28 × 10–3 T – 4.028 × 10–6 T2 ... J/(mol·K)
o
Cp for N2 = 27.03 + 5.815 × 10–3 T – 0.289 × 10–6 T2 ... J/(mol·K)
∆H2 = Enthalpy of the product stream at T over 298 K
= H of CO2 + H of O2 + H of N2
T

= ⌠
⌡ 0.20 (26.54 + 42.45 × 10–3 T – 14.298 × 10–6 T2) dT
298
T

+⌠
⌡ 1.5524 (27.03 + 5.815 × 10–3 T – 0.289 × 10–6 T2) dT
298
T

+⌠
⌡ 0.10 (25.61 + 13.26 × 10–3 T – 4.028 × 10–6 T2) dT
298
T

∆H2 = ⌠
⌡ (49.8304 + 18.8432 × 10–3 T – 3.711 × 10–6 T2) dT
298

18.8432 × 10–3 2
= 49.8334 (T – 298) + 2 T (
2 – 298 )
3.711 × 10–6 3
– 3 T (
3 – 298 )
2
= 49.8334 (T – 298) + 9.4216 × 10–3 T2 – 298 ( )
3
(
– 1.237 × 10–6 T3 – 298 )
We have, ∆H1 = 0
o
Net enthalpy change = ∆H = ∆H2 + Total ∆H298 – ∆H1
o
= ∆H2 + Total ∆H298 – 0
o
= ∆H2 + Total ∆H298
For an adiabatic reaction,
∆H = 0
o
∴ 0 = ∆H2 + Total ∆H298
o
– Total ∆H298 = ∆H2
o
Substituting for total ∆H298 and ∆H2 gives
2
– (– 56635.6) = 49.8334 (T – 298) + 9.4216 × 10–3 T2 – 298 ( )
3
(
– 1.237 × 10–6 T3 – 298 )
Chemical Engineering Thermodynamics - I 3.54 P-V-T Behaviour and Heat Effects

2
(
56635.6 = 49.8334 (T – 298) + 9.4216 × 10–3 T2 – 298 )
3
(
– 1.237 × 10–6 T3 – 298 )
Evaluate the value of T by a trial and error procedure.
Put a value of T, evaluate the R.H.S. and check for L.H.S. ≈ R.H.S. If not, then repeat the
procedure till we get R.H.S. ≈ L.H.S.
Put T = 1000 K
– 2  2
For T = 1000 K, R.H.S. = 49.8334 (1000 – 298) + 9.4216 × 10–3 1000 – 298 ( )
– 3  3
(
– 1.237 × 10–6 1000 – 298 )
= 42363.7
L.H.S. (56635.6) ≠ R.H.S. (42363.7)
For T = 1100 K, R.H.S. = 48916.14 L.H.S. = 56635.6
For T = 1200 K, R.H.S. = 55575.36 L.H.S. = 56635.6
For T = 1210 K, R.H.S. = 56246.87 L.H.S. = 56635.6
For T = 1215 K, R.H.S. = 56683 L.H.S. = 56635.6
R.H.S. (56683) ≈ L.H.S. (56635.6) [Difference is 0.08% only] for T = 1215 K
∴ The theoretical flame temperature is 1215 K ... Ans.
Example 3.11 :
Methane is burned with air. Initially both are at 298 K. The flame temperature is 1598 K.
The combustion reaction is
CH4 + 2 O2 –→ CO2 + 2 H2O
The standard heat of reaction at 298 K is – 8.028 × 105 J/mol methane burned. The mean
heat capacities in J/(mol·K) are 51.66 for CO2, 40.45 for H2O, 34.01 for O2 and 32.21 for N2.
Assuming complete combustion, determine the excess air used.
Solution :
Basis : 1 mol of CH4 charged.
CH4 + 2 O2 –→ CO2 + 2 H2O
Given : Complete combustion means 100% conversion of CH4.
Moles of CH4 reacted = Moles of CH4 charged = 1 mol
1 mol CH4 + 2 mol O2
Theoretical requirement of O2 = 2 × 1 = 2 mol
 for 1 mol CH4 charged  1
Chemical Engineering Thermodynamics - I 3.55 P-V-T Behaviour and Heat Effects

Let x be the moles of O2 in the supplied air.


1
O2 reacted = 1 × 2 = 2 mol

O2 unreacted = (x – 2) mol
79
N2 in the supplied air = 21 · x = 3.762 x mol

1
CO2 produced = 1 × 1 = 1 mol

2
H2O produced = 1 × 1 = 2 mol

Reactant stream : 1 mol CH4 + x mol O2 + 3.762 x mol N2 (i.e., 4.762 x mol air)
Both reactants are at 298 K.
∆H1 = Enthalpy of reactants (reactant stream)
at 298 K over 298 K
= 0 as T of reactants and base T are equal
The standard heat of reaction at 298 K is
o
∆H298 = – 8.028 × 105 J/mol CH4 reacted
Amount of CH4 reacted = 1 mol
o – 8.028 × 105
Total ∆H298 = 1 × 1 = – 8.028 × 105 J
o
Cpm in J/(mol·K) : CO2 = 50.66, H2O = 40.45, O2 = 34.01 and N2 = 32.21
N2 in the product stream = N2 in the air supplied = 3.762 x mol
… as N2 being inert
Product stream : 1 mol CO2 + 2 mol H2O + 3.762 x mol N2 + (x – 2) mol O2
The flame temperature is 1598 K, so the product is at 1598 K.
∆H2 = Enthalpy of product stream at 1598 K over 298 K
o
= ∑ ni Cpmi (1598 – 298)
= 1 × 51.66 (1598 – 298) + 2 × 40.45 (1598 – 298)
+ 3.762 x · (32.21) (1598 – 298)
= 83902 + 201739.23 x J
∆H = Net enthalpy change = Q
o
= ∆H2 + Total ∆H298 – ∆H1
We have, ∆H1 = 0
o
∴ ∆H = ∆H2 + Total ∆H298
Chemical Engineering Thermodynamics - I 3.56 P-V-T Behaviour and Heat Effects

For an adiabatic reaction, ∆H = Q = 0


o
∴ 0 = ∆H2 + Total ∆H298
o
– Total ∆H298 = ∆H2
– (– 8.028 × 105) = 83902 + 201739.23 x
∴ x = 3.5635 mol
Moles of O2 in the air supplied = 3.5635 mol
Moles of O2 theoretically required = 2 mol
O2 in air supplied – O2 theoretical
% excess air = % excess O2 = O2 theoretical × 100

3.5635 – 2
= 2 × 100

= 78.175 ≈ 78.2 ... Ans.


Example 3.12 :
In a laboratory, a steam boiler is fired with liquefied petroleum gas. It may be treated as
pure n-butane. 100% excess air is used. The fuel and air enter the combustion chamber at 298 K.
The flue gases leave the boiler at 523 K. Determine the amount of energy transferred as heat in
the boiler for 15 kg fuel. Assume complete combustion and insulated boiler.
(i) The standard heat of combustion (net heating value) of n-butane is – 2635.58 kJ/mol.
o
(ii) The constants in the heat capacity equation Cp = a + bT + eT–2, kJ/(kmol·K), are as
given below :
Component a b × 103 e × 10–5
CO2 45.369 8.688 – 9.619
O2 30.255 4.207 – 1.887
H2O (g) 28.850 12.055 1.066
N2 27.270 4.930 0.333
Solution :
Basis : 15 kg fuel [n-butane (C4H10)] charged.
Mol. wt. of n-butane = 58
15
Moles of n-butane charged = 58 = 0.2586 kmol

C4H10 + 6.5 O2 → 4 CO2 + 5 H2O (g)


Given : Complete combustion, so all butane is converted.
Butane reacted = Butane fed = 0.2586 kmol
1 kmol C4H10 ≡ 6.5 kmol O2 + 4 kmol CO2 + 5 kmol H2O
Chemical Engineering Thermodynamics - I 3.57 P-V-T Behaviour and Heat Effects

4
CO2 produced = 1 × 0.2586 = 1.0344 kmol

5
H2O produced = 1 × 0.2586 = 1.293 kmol

6.5
O2 reacted = 1 × 0.2586 = 1.6809 kmol

6.5
Theoretical O2 required for 0.2586 kmol C4H10 fed = 1 × 0.2586 = 1.6809 kmol

Excess air used = 100%


Excess O2 = Excess air = 100%

O2 in the air supplied = O2 theoretically required 1 +


% excess
 100 

= 1.6809 1 + 100 = 3.3618 kmol


100
 
79
N2 in the air supplied = 21 × 3.3618 = 12.647 kmol

Material balance of O2 :
O2 unreacted = O2 supplied – O2 reacted
= 3.3618 – 1.6809 = 1.6809 kmol
N2 in the product stream = N2 in the air supplied ... as N2 being inert
= 12.647 kmol
The standard heat of reaction at 298 K is – 2635.58 kJ/mol … Given
Amount of butane reacted = 0.2586 kmol = 258.6 mol
o – 2635.58
∴ Total ∆H298 = 1 × 258.6 = – 681560.99 kJ
Reactant stream : 0.2586 kmol C4H10 + 3.3618 kmol O2 + 12.647 kmol N2
T = 298 K
∆H1 = Enthalpy of reactant stream at 298 K over 298 K
= 0 ... as T of reactant stream = Base temperature (298 K)
Product stream : 1.6809 kmol O2 + 12.647 kmol N2 + 1.293 kmol H2O + 1.0344 kmol CO2
T = 523 K
o
Cp (CO2) = 45.369 + 8.688 × 10–3 T – 9.619 × 105 T–2
o
Cp (O2) = 30.255 + 4.207 × 10–3 T – 1.887 × 105 T–2
o
Cp (H2O) = 28.850 + 12.055 × 10–3 T + 1.066 × 105 T–2
o
Cp (N2) = 27.270 + 4.930 × 10–3 T + 0.333 × 105 T–2
o
Cp = a + bT + eT–2
Chemical Engineering Thermodynamics - I 3.58 P-V-T Behaviour and Heat Effects

∆H2 = Enthalpy of product stream at 523 K over 298 K


523

= ⌠
⌡ Σ niCpi dT
298

where ni is the number of moles of ith component.


o
For Cp = a + bT + eT–2
  2 1 
∆H of ith component = ni a (T – 298) + 2 T2 – 298 – e T – 298
b 1

( 
) 
523

∆H2 = ⌠
⌡ 1.0344 (45.369 + 8.688 × 10–3 T – 9.619 × 105 T–2) dT
298
523

+⌠
⌡ 1.6809 (30.255 + 4.207 × 10–3 T – 1.887 × 105 T–2) dT
298
523

+⌠
⌡ 1.293 (28.850 + 12.055 × 10–3 T + 1.086 × 105 T–2) dT
298
523

+⌠
⌡ 12.467 (27.270 + 4.930 × 10–3 T + 0.333 × 105 T–2) dT
298
523

∆H2 = ⌠
⌡ [475.063 + 93.1078 × 10–3 T – 7.5893 × 105 × T–2] dT
298

93.1078 × 10–3  2  2
= 475.063 (523 – 298) + 2
[
523 – 298 ]
+ 7.5893 × 105 523 – 298
1 1
 
= 106889.17 + 8599.67 – 1095.63
= 114393.21 kJ
Note that ⌠ T–2 dT = – 1 
 ⌡ T
Energy released is used only to increase T of product stream ... insulated boiler
∆H = Net enthalpy change = Q = Energy transferred as heat
o
= ∆H2 + Total ∆H298 – ∆H1
o
= ∆H2 + Total ∆H298 – 0 ... since ∆H1 = 0
= 114393.21 + (– 681560.99) – 0
= – 567167 kJ = – 567.167 MJ ≈ – 567.17 MJ
The minus (–ve) sign indicates that the heat is removed.
Energy transferred as heat (heat removed) = 567.17 MJ ... Ans.
Chemical Engineering Thermodynamics - I 3.59 P-V-T Behaviour and Heat Effects

Example 3.13 :
Propane is burned with air and the resulting hot gases are used in the boiler to raise the
steam. The flue gas leaving the boiler at 550 K has the following composition on dry basis :
CO2 = 6.28%, CO = 3.14%, O2 = 7.85% and 82.73% (by volume).
Propane and air enter the combustion chamber at 298 K. Determine the amount of energy
transferred as heat in the boiler per kg of propane burned.
o o
Data : ∆Hf (C3H8) = – 103.833 kJ/mol, ∆Hf 298 (CO) = – 110.532 kJ/mol,
o o
∆Hf 298 (CO2) = – 393.978 kJ/mol and ∆Hf 298 [H2O (g)] = – 241.997 kJ/mol
o
Cp data are as follows :
o
Cp = a + bT + eT–2 in J/(mol·K)
Component a b × 103 e × 10–5
CO2 45.369 8.688 – 9.619
O2 30.255 4.207 – 1.887
CO 28.068 4.631 – 0.258
N2 27.270 4.930 0.333
H2O (g) 28.850 12.055 1.006
Assume complete combustion and insulated boiler.
Solution :
Basis : 100 mol dry flue gas.
Moles of CO2 = 0.0628 × 100 = 6.28 mol

Moles of CO = 0.0314 × 100 = 3.14 mol


Similarly, Moles of O2 = 7.85 mol
Moles of N2 = 82.73 mol
(i) C3H8 + 5 O2 –→ 3 CO2 + 4 H2O (g)
(ii) C3H8 + 3.5 O2 –→ 3 CO + 4 H2O (g)
As N2 being an inert gas,
N2 in the flue gas = N2 in the air supplied
∴ N2 in the air supplied = 82.73 mol
Air contains 79 mole % N2 and 21 mole % O2.
21
O2 in the air supplied = 79 × 82.73 = 21.99 mol
Chemical Engineering Thermodynamics - I 3.60 P-V-T Behaviour and Heat Effects

We have, 5 mol O2 ≡ 3 mol CO2 ... Reaction (i)


5
O2 reacted by (i) = 3 × 6.28 = 10.467 mol

[6.28 mol CO2 is produced by reaction (i)]


CO produced by reaction (ii) = 3.14 mol
We have, 3.5 mol O2 ≡ 3 mol CO ... Reaction (ii)
3.5
O2 reacted by (ii) = 3 × 3.14 = 3.663 mol

O2 totally reacted = 10.467 + 3.663 = 14.13 mol


O2 in the flue gas = O2 unreacted = 7.85 mol
O2 in the air = O2 reacted + O2 unreacted
21.99 = 14.13 + 7.85 = 21.98 ... Calculations are correct
By reaction (i), 3 mol CO2 ≡ 4 mol H2O
4
H2O produced by (i) = 3 × 6.28 = 8.373 mol

By reaction (ii) : 3 mol CO + 4 mol H2O


4
H2O produced by (ii) = 3 × 3.14 = 4.187 mol

Total H2O produced = 8.373 + 4.187 = 12.56 mol


1 mol C3H8 + 3 mol CO2 … by reaction (i)
1
∴ C3H8 reacted by (i) = 3 × 6.28 = 2.093 mol

1 mol C3H8 + 3 mol CO … by reaction (ii)


1
∴ C3H3 reacted by (ii) = 3 × 3.14 = 1.0466 mol

Total C3H8 reacted = 2.0933 + 1.0466 = 3.1399 mol


Combustion is complete.
∴ C3H8 charged = C3H8 reacted = 3.1399 mol
We will calculate the amount of energy transferred as heat in the boiler based on
3.1399 mol C3H8 and then we will convert it based on 1 kg C3H8 fed and burned.
Reactant stream : 3.1399 mol C3H8 + 21.99 mol O2 + 82.73 mol N2,
T = 298 K , Base T = 298 K
∆H1 = Enthalpy of the reactant stream at 298 K over 298 K
= 0 ... as T of the reactant stream = base temperature
Chemical Engineering Thermodynamics - I 3.61 P-V-T Behaviour and Heat Effects

Let us determine the total standard heat of reaction at 298 K.


(i) C3H8 + 5 O2 –→ 3 CO2 + 4 H2O (g)
o o o o
∆H298 (i) = 3 × ∆Hf 298 [CO2] + 4 × ∆Hf 298 [H2O] – 1 × ∆Hf 298 [C3H8]
o
– 5 × ∆Hf 298 [O2]
= 3 × (–393.978) + 4 × (–241.997) – 1 × (– 103.833) – 5 × (0)
= – 2046.089 kJ/mol C3H8
(ii) C3H8 + 3.5 O2 –→ 3 CO + 4 H2O (g)
o
∆H298 (ii) = 3 × (–110.532) + 4 × (–241.997) – 1 × (–103.833) – 3.5 × (0)
= – 1195.75 kJ/mol C3H8
C3H8 burned by reaction (i) = 2.093 mol
C3H8 burned by reaction (ii) = 1.0466 mol
Total standard heat of reaction of the above reactions at 298 K is
o o
o ∆H298 (i) ∆H298 (ii)
Total ∆H298 = 1 × 2.093 + 1 × 1.0466

– 2046.089 – 1195.75
= 1 × 2.093 + 1 × 1.0466

= – 5533.936 kJ
= – 5533.94 × 103 J
Product stream : 82.73 mol N2 + 7.85 mol O2 + 6.28 mol CO2 + 3.14 mol CO + 12.56 mol
H2O.
T = 550 K
o
Cp (CO2) = 45.369 + 8.688 × 10–3 T – 9.619 × 105 T–2
o
Cp (CO) = 28.068 + 4.631 × 10–3 T – 0.258 × 105 T–2
o
Cp (O2) = 30.255 + 4.207 × 10–3 T – 1.887 × 105 T–2
o
Cp (N2) = 27.270 + 4.930 × 10–3 T + 0.333 × 105 T–2
o
Cp (H2O) = 28.850 + 12.055 × 10–3 T + 1.066 × 105 T–2
∆H2 = Enthalpy of the product stream (flue gas) at 550 K over 298 K
550

= ⌠
o
⌡ ni Cpi dT ni - moles of ith component
298
Chemical Engineering Thermodynamics - I 3.62 P-V-T Behaviour and Heat Effects

550

= ⌠
⌡ 6.28 (45.269 + 8.688 × 10–3 T – 9.619 × 105 T–2) dT
298
550

+⌠
⌡ 3.14 (28.068 + 4.631 × 10–3 T – 0.258 × 105 T–2) dT
298
550

+⌠
⌡ 7.85 (30.255 + 4.207 × 10–3 T – 1.887 × 105 T–2) dT
298
550

+⌠
⌡ 82.73 (27.270 + 4.930 × 10–3 T + 0.333 × 105 T–2) dT
298
550

+⌠
⌡ 12.56 (28.850 + 12.055 × 10–3 T + 1.066 × 105 T–2) dT
298
550

∆H2 = ⌠
⌡ (3228.956 + 661.397 × 10–3 T – 35.092 × 105 T–2) dT
298

661.397 × 10–3  2  2
= 3228.956 (550 – 298) + 2
(550 – 298 )
+ 35.092 × 105 550 – 298
1 1
 
= 813696.91 + 70668.95 – 5395.47
= 878970.39 J
∆H = Net enthalpy change = Energy transferred as heat
o
= ∆H2 + Total ∆H298 – ∆H1
= 878970.39 + (– 5533.94 × 103) – 0
= – 4654969.61 J ... based on 2.1399 mol C3H8 burned
We have to find out the energy transferred per kg of C3H8 burned.
Actual amount of C3H8 burned = 1 kg
Molecular weight of C3H8 = 44 kg/kmol
1
Moles of actual C3H8 burned = 44 = 0.02273 kmol = 22.73 mol

Hence, energy transferred as heat in the boiler (i.e., heat removed in the boiler)
4654969.61
= 3.1399 × 22.73 = 33697716.24 J

= 33697.71624 kJ = 33.69771624 MJ ≈ 33.70 MJ ... Ans.


Chemical Engineering Thermodynamics - I 3.63 P-V-T Behaviour and Heat Effects

Example 3.14 :
Calculate the standard heat of reaction for the following reaction :
11
2 FeS2 (s) + 2 O2 (g) –→ Fe2O3 (s) + 4 SO2 (g)

The standard heats of formation at 298 K of the components are :


o
Component ∆Hf 298 , kJ/mol
FeS2 (s) – 178.02
Fe2SO3 (s) – 822.71
SO2 (g) – 297.10
Solution :
11
2 FeS2 (s) + 2 O2 (g) –→ Fe2O3 (s) + 4 SO2 (g)

The standard heat of reaction at 298 K is given by


o
∆H298 = Σ [ni ∆Hof 298 (i) ]products – Σ [ni ∆Hof 298 (i) ]reactants
where ni is the stoichiometric coefficient of ith component
o o o
∆H298 = [1 × ∆H f 298 (Fe2SO3) + 4 × ∆Hf 298 (SO2) ]

– 2 × ∆Hf 298 (FeS2) + 2 × ∆Hf 298 (O2) 


o 11 o
 

= [1 × (– 822.71) + 4 × (– 297.10)] – 2 × (– 178.02) + 2 (0)


11
 
= – 1655.07 kJ/mol FeS2 reacted or – 1655.07 kJ ... Ans.
Example 3.15 :
The heat of reaction at 300 K and 1 atm for A (g) + 3B (g) –→ C (g) is – 209.35 kJ/mol A
reacted. The heat capacities of the components in J/(mol·K) are :
o
Cp for A = – 1.6748 + 33.496 × 10–2 T where T is in K,
o
Cp for B = 29.309 and Cp for C = 108.862
Calculate the heat of reaction at 500 K and at 1 atm.
Solution :
A (g) + 3B (g) –→ C (g)
o
Given : ∆H300 = – 209.35 kJ/mol A reacted = – 209.35 × 103 J/mol
o
Cp for A = – 1.6748 + 33.496 × 10–2 T
o
Cp for B = 29.309
o
Cp for C = 108.862 J/(mol·K)
Chemical Engineering Thermodynamics - I 3.64 P-V-T Behaviour and Heat Effects

o
For Cp = a + bT, the standard heat of reaction at T is given by
o ∆b
∆HT = ∆Ho + ∆a T + 2 T2

∆a = Σ (niai)products – Σ (niai)reactants
ni is the stoichiometric coefficient of ith component.
∆a = [1 × 108.862] – [1 × (– 1.6748) + 3 × 29.309]
= 22.6098
∆b = [1 × 0] – [1 × 33.496 × 10–2 + 3 × (0)] = – 33.496 × 10–2
o 33.496 × 10–2 2
∆HT = ∆Ho + 22.6098 T – 2 T
o
∆HT = ∆Ho + 22.6098 T – 16.748 × 10–2 T2 ... in J/mol
o o
To evaluate the value of ∆Ho, put T = 300 K and ∆HT = ∆H300
– 209.35 × 103 = ∆Ho + 22.6098 (300) – 16.748 × 10–2 (300)
∆Ho = – 201059.74 J/mol
o
∆HT = – 201059.74 + 22.6098 T – 16.748 × 10–2 T2
Evaluation of the heat of reaction at 500 K :
o
∆H500 = – 201059.74 + 22.6098 (500) – 16.748 × 10–2 (500)2
= – 231624.84 J/mol
= – 231.62484 ≈ – 231.625 kJ/mol ... Ans.
Example 3.16 :
Carbon monoxide at 1000 K (727°C) is burned with air at 800 K (527°C). 90% excess air is
used. The products of combustion leave the reaction chamber at 1250 K (977°C). The standard
heat of reaction at 298 K is – 283.028 kJ/mol CO burned. The mean heat capacities applicable in
the temperature range of this problem for CO, CO2, O2 and N2 are 29.38, 49.91, 33.13 and
31.43 J/(mol·K) respectively. Calculate the heat evolved in the reaction chamber per kmol of CO
burned.
Solution :
Basis : 1 kmol of CO burned.
1
CO + 2 O2 –→ CO2
Moles of CO2 produced from 1 kmol CO burned = 1 kmol
1/2
Moles of O2 reacted = 1 × 1 = 0.5 kmol
1/2
Theoretical O2 requirement for 1 kmol CO fed = 1 × 1 = 0.5 kmol
Chemical Engineering Thermodynamics - I 3.65 P-V-T Behaviour and Heat Effects

[Whatever CO fed is CO burned.]


% excess air = 90
% excess O2 = % excess air = 90

O2 in the supplied air = 1 + 100  × O2 theoretically required


% excess
 

= 1 + 100 × 0.5 = 0.95 kmol


90
 
79
N2 in the air supplied = 21 × 0.95 = 3.574 kmol

[Air contains 21% O2 and 79% N2 by volume.]


O2 unreacted = O2 in air supplied – O2 reacted
= 0.95 – 0.50 = 0.45 kmol
Input to the reactor : 1 kmol CO + 0.95 kmol O2 + 3.574 kmol N2
Output from the reactor : 0.45 kmol O2 + 3.574 kmol N2 + 1 kmol CO2

Fig. E 3.16
∆H1 = Enthalpy of the reactant stream at T over 298 K
= Enthalpy of CO at 1000 K over 298 K + Enthalpy of O2 and N2 at 800 K over 298 K
o
Cpm : For CO = 29.38, for O2 = 33.13, for N2 = 31.43 J/(mol·K)
Moles of CO = 1 kmol = 1000 mol
Moles of O2 = 0.95 kmol = 950 mol
and Moles of N2 = 3574 mol
o
∆H1 = ∑ ni Cpm ∆T
∴ ∆H1 = 1000 × 29.38 (1000 – 298) + 950 × 33.13 (800 – 298) + 3574 × 31.43 (800 – 298)
= 9.281 × 107 J
o
∆H298 = – 283.028 kJ/mol CO burned = – 283028 J/mol CO burned
Moles of CO burned = 1 kmol = 1000 mol
o – 283028
Total ∆H298 = 1 × 1000 = – 283028 × 103 J

∆H2 = Enthalpy of product stream at 1250 K over 298 K


Chemical Engineering Thermodynamics - I 3.66 P-V-T Behaviour and Heat Effects

Product stream : 1 kmol = 1000 mol CO2, 3574 mol N2, 0.45 kmol = 450 mol O2
o
Cpm : for CO2 = 49.91, for N2 = 31.43 and for O2 = 33.13 J/(mol·K)
o
∆H2 = ∑ ni CP mi ∆T
∴ ∆H2 = 1000 × 49.91 (1250 – 298) + 450 × 33.13 (1250 – 298) + 3574 × 31.43 (1250 – 298)
= (1000 × 49.91 + 450 × 33.13 + 3574 × 31.43) (1250 – 298)
= 1.686 × 108 J
Net enthalpy change = ∆H = Q
o
∆H = ∆H2 + Total ∆H298 – ∆H1
= 1.686 × 108 + (– 283028 × 103) – 9.281 × 107
= – 207238000 J = – 207.238 × 103 kJ
Where the minus (–ve) sign indicates that the heat is evolved during the reaction.
Heat evolved in the reaction chamber per kmol CO burned = 207.238 × 103 kJ ... Ans.
Example 3.17 :
Calculate the standard heat of reaction at 800 K for the complete combustion of pentane gas
[C5H12]. The mean heat capacities of C5H12, O2, CO2 and H2O are 247, 33.62, 52.32 and
38.49 J/(mol·K), respectively. The standard heat of combustion at 298 K is – 3271.71 kJ/mol.
Solution :
C5H12 + 8 O2 –→ 5 CO2 + 6 H2O
Given : Standard heat of combustion of pentane gas at 298 K is – 3271.71 kJ/mol
o o
∴ ∆H298 = ∆Hc 298 = – 3171.71 kJ/mol = – 3271.71 × 103 J/mol
o
Cpm for C5H12 = 247 Cpm for O2 = 33.62
o
Cpm for CO2 = 52.32 and Cpm for H2O = 38.49 J/(mol·K)
o
For the Cpm data of the components provided, the standard heat of reaction at any
temperature T is given by
o o
∆HT = ∆Ho + ∆Cpm T
o
∆Cpm = Σ (niCopmi )products – Σ (niCopmi )reactants
where ni is the stoichiometric coefficient of ith component.
n for C5H12 = 1, n for O2 = 8, n for CO2 = 5 and n for H2O = 6
o
∆Cpm = [5 × 52.32 + 6 × 38.49] – [1 × 247 + 8 × 33.62]
= – 23.42
o
∆HT = ∆Ho – 23.42 T ... J/mol
Chemical Engineering Thermodynamics - I 3.67 P-V-T Behaviour and Heat Effects

o o
To evaluate the value of ∆Ho substitute T = 298 K and ∆HT = ∆H298 in the above equation.
o o
Cpm data is in J/(mol·K), so use ∆H298 in J/mol.
o
∆H298 = – 3171.71 × 103 J/mol
– 3171.71 × 103 = ∆Ho – 23.42 (298)
∴ ∆Ho = – 3264.73 × 103 J/mol
o
The desired relationship between ∆HT and T is
o
∆HT = – 3264.73 × 103 – 23.42 T ... in J/mol
o
To find ∆HT at T = 800 K :
o
∆H800 = – 3264.73 × 103 – 23.42 (800)
= – 3283466 J/mol = 3283.47 kJ/mol ... Ans.
Example 3.18 :
Methanol is produced according to the following reaction :
CO (g) + 2H2 (g) –→ CH3OH (g)
The standard heats of formation at 298 K for CO and CH3OH are – 110.125 kJ/mol and
– 200.660 kJ/mol, respectively.
The heat capacity data in J/(mol·K) for the reaction participants are :
o
Cp (CH3OH) = 19.382 + 101.564 × 10–3 T – 28.683 × 10–6 T2
o
Cp (CO) = 28.068 + 4.631 × 10–3 T – 2.5773 × 104 T–2
o
Cp (H2) = 27.012 + 3.509 × 10–3 T + 6.9006 × 104 T–2
(i) Express the heat of reaction as a function of temperature.
(ii) Calculate the heat of reaction at 1073 K.
Solution :
First, we will find the standard heat of reaction at 298 K and then we will make use of this
o
value to evaluate ∆Ho of the desired expression of ∆HT .
CO (g) + 2 H2 (g) –→ CH3OH (g)
o
∆Hf 298 for CO = – 110.125 kJ/mol
o
∆Hf 298 for CH3OH = – 200.660 kJ/mol
The standard heat of reaction at 298 K is given by
o
∆H298 = Σ (∆Hof 298 )products – Σ (∆Hof 298 )reactants
o o o
= (1 × ∆Hf 298 )CH – (1 × ∆Hf 298 )CO – (2 × ∆Hf 298 )H
3OH 2

= 1 × (– 200.660) – 1 × (– 110.125) – 2 × (0)


= – 90.535 kJ/mol = – 90535 J/mol of CO reacted
Chemical Engineering Thermodynamics - I 3.68 P-V-T Behaviour and Heat Effects

o
From the data provided, Cp can be expressed as
o
Cp = a + bT + cT2 + eT–2 , J/(mol·K)
Let us evaluate ∆a, ∆b, ......
∆a = Σ (niai)products – Σ (niai)reactants
where ni is the stoichiometric coefficient of ith component participating in a given chemical
reaction.
∆a = [1 × 19.382] – [1 × 28.068 + 2 × 27.012] = – 62.71
Similarly,
∆b × 103 = 1 × 101.564 – 1 × 4.631 – 2 × 3.503 = 89.927
∴ ∆b = 89.927 × 10–3
∆c × 106 = 1 × (– 28.683) – 0 = – 28.863 ... no c values are provided for CO and H2
∴ ∆c = – 28.683 × 10–6
∆e × 10–4 = 1 (0) – 1 × (– 2.5773) – 2 × (6.9006)
= – 11.2339 ... no d value is provided for CH3OH
∴ ∆e = – 11.2339 × 104
The standard heat of reaction at any temperature T for Cp data of the type
Cp = a + bt + cT2 + eT–2 is given by
∆b ∆c
∆HT = ∆Ho + ∆a T + 2 T2 + 3 T3 – ∆e T
o 1
... (A)
 
as ⌠ T–2 dT = – 1 
 ⌡ T
Substituting the values of ∆a, ∆b, ...... yields
89.927 × 10–3 2 28.683 × 10–6 3
T + 11.2239 × 104 T ... in J/mol
o 1
∆HT = ∆Ho – 62.71 T + T –
2 3  
[Every term of the above equation is in J/mol as Cp is in J/(mol·K)]

∆HT = ∆Ho – 62.71 T + 44.9635 × 10–3 T2 – 9.561 × 10–6 T3 + 11.2239 × 104 T
o 1
… (B)
 
o o
To determine the value of ∆Ho, substitute T = 298 K and ∆HT = ∆H298
o
We have, ∆H298 = – 90535 J/mol
∴ – 90535 = ∆Ho – 62.71 (298) + 44.9635 × 10–3 (298)2

– 9.561 × 10–6 (298)3 + 11.2239 × 104 298


1
 
∆Ho = – 75960 J/mol
Chemical Engineering Thermodynamics - I 3.69 P-V-T Behaviour and Heat Effects

Substituting this value of ∆Ho in Equation (B), we get the relation for the heat of reaction as
a function of temperature for the given reaction.

∆HT = – 75960 – 62.71 T + 44.9635 × 10–3 T2 – 9.561 × 10–6 T3 + 11.2239 × 104 T
o 1
 
... in J/mol ... Ans. (i)
To calculate the heat of reaction at 1073 K : For this substitute T = 1073 K in the above
expression.
o 11.2239 × 104
∆H1073 = –75960 – 62.71 (1073) + 44.9635 × 10–3 (1073)2 – 9.561 × 10–6 (1073)3 + 1073
= – 103186.87 = – 103187 J/mol = – 103.187 kJ/mol
Heat of reaction at 1073 K = – 103.187 kJ/mol ... Ans. (ii)
o
(As ∆HT is –ve, the reaction is exothermic.)
Example 3.19 :
Ammonia is synthesized as per the following reaction :
1 3
2 N 2 + 2 H2 –→ NH3
The standard heat of reaction at 298 K for this reaction is – 46.222 kJ.
The heat capacities of the reaction participants are represented by
o
Cp = α + β T + γT2
o
where Cp is in J/(mol·K) and the constants are :
Component α β γ
NH3 25.48 36.89 × 10–3 – 6.305 × 10–6
N2 27.31 5.2335 × 10–3 – 4.1868 × 10–9
H2 29.09 – 8.374 × 10–4 2.0139 × 10–6
Obtain an expression relating the heat of reaction and the temperature of the reaction.
Solution :
1 3
2 N2 + 2 H2 –→ NH3
o
∆H298 = – 46.222 kJ = – 46222 J/mol NH3 formed
o o
For Cp data represented as Cp = α + βT + γT2, the standard heat of reaction at any
temperature T is given by
o
∆HT = ∆Ho + ∆α T + ∆β T2 + ∆γ T3 ... in J/mol
[As Cp data are given in J/(mol·K)]
Chemical Engineering Thermodynamics - I 3.70 P-V-T Behaviour and Heat Effects

Let us evaluate ∆α, ∆β and ∆γ.


∆α = Σ (niαi)products – Σ (niαi)reactants
where ni is the stoichiometric coefficient of ith component.
1 3
n for NH3 = 1, n for N2 = 2 and n for H2 = 2

∆α = [1 × 25.48] – 2 × (27.31) + 2 × (29.09) = – 31.81


1 3

 
β for H2 = – 8.374 × 10 = – 0.8374 × 10
–4 –3

∆β × 103 = [1 × (36.89)] – 2 (5.2335) + 2 (– 0.8374) = 35.529


1 3
 
∴ ∆β = 35.529 × 10–3
γ for N2 = – 4.1868 × 10–9 = (– 4.1868 × 10–3) × 10–6

∆γ × 106 = [1 × (– 6.305)] – 2 (– 4.1868 × 10–3) + 2 (2.0139) = – 9.3237


1 3
 
∴ ∆γ = – 9.3237 × 10–6
Substituting the numerical values of ∆α, ∆β and ∆γ yields,
o 35.529 × 10–3 2 9.3237 × 10–6 3
∆HT = ∆Ho – 31.81 T + 2 T – 3 T
o
∆HT = ∆Ho – 31.81 T + 17.7647 × 10–3 T2 – 3.1079 × 10–6 T3 ... in J/mol
o
Given : ∆H298 = – 46.222 kJ/mol = – 46222 J/mol NH3 formed
o o
To determine the value of ∆Ho , substitute T = 298 K and ∆HT = ∆H298
– 46222 = ∆Ho – 31.81 (298) + 17.7647 × 10–3 (298)2 – 3.1079 × 10–6 (298)3
∴ ∆Ho = – 38237.95 ≈ – 38238 J/mol
∴ The desired expression for the heat of reaction as a function of temperature is
o
∆HT = – 38238 – 31.81 T + 17.1995 × 10–3 T2 – 3.1079 × 10–6 T3 ... in J/mol NH3 ... Ans.
Example 3.20 :
Pure CO is burned with 100% excess air. Combustion is 80% complete. The reactants are at
373 K and the products are at 573 K. Using the following data, calculate the quantity of heat
added or removed per kmol of CO fed. Mean molar heat capacity data in kJ/(kmol·K) are :
o o
Component Cpm (373 – 298 K) Cpm (573 – 298 K)
CO 29.22 30.61
CO2 – 43.77
O2 29.84 30.99
N2 29.17 29.66
o o
∆Hf 298 (CO) = – 110524 kJ/kmol and ∆Hf 298 (CO2) = – 393514 kJ/kmol
Chemical Engineering Thermodynamics - I 3.71 P-V-T Behaviour and Heat Effects

Solution :
Basis : 1 kmol of CO fed.
Combustion is 80% complete. So,
Moles of CO reacted = 0.80 × 1 = 0.8 kmol
1
CO + 2 O2 → CO2

1/2
Theoretical O2 requirement for 1 kmol CO fed = 1 × 1 = 0.5 kmol

% excess air = 100


% excess O2 = % excess air = 100

O2 in supplied air = Theoretical O2 1 + 100 


% excess
 

= 0.5 1 + 100 = 1 kmol


100
 
79
N2 in supplied air = 21 × 1 = 3.762 kmol

1/2
O2 reacted for 0.80 kmol CO reacted = 1 × 0.80 = 0.40 kmol

O2 unreacted = 1 – 0.40 = 0.6 kmol


1
CO2 produced = 1 × 0.80 = 0.8 kmol

N2 in product stream = N2 in air supplied ... as N2 being inert


= 3.762 kmol
CO unreacted = 1 – 0.80 = 0.20 kmol
Reactant stream : 1 kmol CO + 1 kmol O2 + 3.762 kmol N2
T = 373 K
o
Cpm : For CO = 29.22, for O2 = 29.84 and for N2 = 29.17 kJ/(kmol·K)
o
∆H1 = ∑ ni Cpmi ∆T = Enthalpy of reactant stream at 373 K over 298 K
= [1 × 29.22 + 1 × 29.84 + 3.762 × 29.17] [373 – 298]
= 12659.81 kJ
1
CO + 2 O2 –→ CO2
o
∆Hf 298 (CO) = – 110524 kJ/kmol
o
∆Hf 298 (CO2) = – 393514 kJ/kmol
Chemical Engineering Thermodynamics - I 3.72 P-V-T Behaviour and Heat Effects

The standard heat of reaction at 298 K for this reaction is given by


o o o 1 o
∆H298 = 1 × ∆Hf 298 [CO2] – ∆Hf 298 [CO] – 2 ∆Hf 298 [O2]

1
= 1 × (– 393514) – 1 × (– 110524) – 2 (0)

= – 282990 kJ/kmol CO reacted


Amount of CO reacted = 0.80 kmol
Total heat of reaction at 298 K is
o – 282990
Total ∆H298 = 1 × 0.80 = – 226392 kJ

Product stream : 0.2 kmol CO + 0.80 kmol CO2 + 0.60 kmol O2 + 3.762 kmol N2
T = 573 K
o
Cpm : For CO = 30.61, for CO2 = 43.77, for O2 = 30.99 and for N2 = 29.66 kJ/(kmol·K)
∆H2 = Enthalpy of product stream at 573 K over 298 K
= [0.20 × 30.61 + 0.80 × 43.77 + 0.60 × 30.99 + 3.762 × 29.66) (573 – 298)
= 47111.05 kJ
Net enthalpy change = ∆H = Q
o
= ∆H2 + Total ∆H298 – ∆H1
= 47111.05 + (– 226392) – 12659.81
= – 191940.76 kJ
The minus sign indicates that the heat is evolved in the reaction.
Heat evolved = 191940.76 kJ ... Ans.
Example 3.21 :
Carbon monoxide reacts with water vapour to produce carbon dioxide and hydrogen :
CO (g) + H2O (g) –→ CO2 (g) + H2 (g)
The standard heat of reaction of this reaction at 298 K is – 41.190 kJ.
The reactants are used in stoichiometric proportions and are at 298 K. 75% of CO is
converted by the above reaction. The products leave the reactor at 800 K. Calculate the amount
of heat to be added or removed in the reactor per 1000 kg of hydrogen production.
The mean heat capacities in J/(mol·K) of the components are :
CO = 30.35, CO2 = 45.64, H2O (vapour) = 36 and H2 = 29.30
Solution :
Basis : 1000 kg of hydrogen produced.
Molecular weight of H2 = 2
Chemical Engineering Thermodynamics - I 3.73 P-V-T Behaviour and Heat Effects

1000
Moles of hydrogen produced = 2 = 500 kmol
CO + H2O –→ CO2 + H2
Conversion of CO = 75%
1 kmol CO ≡ 1 kmol H2
1
Moles of CO reacted to produce 1000 kg H2 = 1 × 500 = 500 kmol

Moles of CO reacted 500


Moles of CO fed = 0.75 = 0.75 = 666.67 kmol

1 kmol CO ≡ 1 kmol H2O


Reactants are in stoichiometric proportions, so
1
H2O fed = 1 × 666.67 = 666.67 kmol

1 kmol CO2 ≡ 1 kmol H2


1
∴ Moles of CO2 produced = 1 × 500 = 500 kmol

Moles of CO unreacted = CO fed – CO reacted


= 666.67 – 500
= 167.67 kmol
Moles of H2O unreacted = 666.67 – 500
= 167.67 kmol
Reactant stream : 500 kmol CO + 500 kmol H2O
Both the reactants are at 298 K.
∆H1 = Enthalpy of reactant stream at 298 K over 298 K
= 0 ... as the reactants are at 298 K and
the reference temperature is also 298 K
o
Given : ∆H298 = – 41.190 kJ/mol H2 produced or CO reacted
Amount of CO reacted = 500 kmol
Total standard heat of reaction at 298 K is
o – 41.190
Total ∆H298 = 1 × 500

= – 20595 kJ = – 20595 × 103 J


Product stream : 166.67 kmol CO + 166.67 kmol H2O + 500 kmol CO2 + 500 kmol H2
Product stream is at 800 K
o
Cpm in J/(mol·K) for CO = 30.35, for CO2 = 45.64, for H2O = 36 and for H2 = 29.30
Chemical Engineering Thermodynamics - I 3.74 P-V-T Behaviour and Heat Effects

o
∆H2 = ∑ niCpmi ∆T = Enthalpy of the product stream at 800 K over 298 K
= HCO + HCO2 + HH2O + HH2

= 166.67 × 30.35 (800 – 298) + 500 × 45.64 (800 – 298)


+ 166.67 × 36 (800 – 298) + 500 × 29.30 (800 – 298)
= 24361334.36 kJ
Net enthalpy change =∆H = Q
o
= ∆H2 + Total ∆H298 – ∆H1
= 24361334.36 – 20595 × 103 – 0
= 3766334.359 kJ
≈ 3.766 × 106 kJ
Where the positive (+ve) sign signifies that the heat is added to the system.
∴ Heat added = 3.766 × 106 kJ ... Ans.
Example 3.22 :
Synthesis gas is manufactured by the catalytic reforming of methane with steam at high T
and atmospheric pressure :
CH4 (g) + H2O (g) → CO (g) + 3 H2 (g) … synthesis gas CO + H2
CO (g) + H2O (g) → CO2 (g) + H2 (g)
The reactants are supplied in the ratio, 2 mol steam to 1 mol CH4. Heat is supplied to the
reactor so that the products react at a temperature of 1300 K and CH4 is completely converted.
The product stream contains 17.4 mole % CO. Assuming that the reactants are at 600 K,
calculate the heat requirement for the reactor.
Data : CH (g) + H O (g) → CO (g) + 3 H (g) ∆H° = 205.813 kJ/mol
4 2 2 298

CO (g) + H2O (g) → CO2 (g) + H2 (g) ° = –41.166 kJ/mol


∆H298
Component ° °
Cpm (298 – 600 K) Cpm (298 – 1300 K)
CH4 44.026 –
H2O 34.826 J/(mol·K) –
CO – 31.702
H2 – 29.994
CO2 – 49.830
H2O – 38.742 J/(mol·K)
Solution : Basis : 1 mol of CH4 and 2 mol of H2O fed to the reactor.
Given : All CH4 reacted.
Chemical Engineering Thermodynamics - I 3.75 P-V-T Behaviour and Heat Effects

∴ Amount of CH4 reacted = 1 mol


CH4 + H2O → CO + 3H2 … (1)
CO + H2O → CO2 + H2 … (2)
Given : Product contains 17.4 mole % CO. Therefore, molefraction of CO = 0.174.
Let x be the mole of CO in the product stream.
Moles of CH4 reacted = 1 mol
1
Moles of H2O reacted by (1) = 1 × 1 = 1 mol
3
Moles of H2 produced by (1) = 1 × 1 = 3 mol
1
Moles of CO produced by (1) = 1 × 1 = 1 mol.
Moles of CO reacted by (2) = CO produced – CO in product stream = (1 – x) mol
1
Moles of H2O reacted by (2) = 1 × (1 – x) = (1 – x) mol
1
Moles of CO2 produced by (2) = 1 (1 – x) = (1 – x) mol
1
Moles of H2 produced by (2) = 1 (1 – x) = (1 – x) mol
Moles of product stream :
CO = x mol = x
CH4 = nil = 0
H2O = 2 – (1 + 1 – x) = x
CO2 = 1 – x = 1–x
H2 = 3 + (1 – x) = 4–x
Total = 5 mol
Mole fraction of CO in the product stream = 0.17
x
5 = 0.174
∴ x = 0.174 × 5 = 0.87 mol
Product stream :
Component mol
CH4 – 0
CO (x) 0.87
H2O (x) 0.87
CO2 (1 – x) 0.13
H2 4–x 3.13
Reactant stream :
CH4 = 1 mol
H2O = 1 mol
Chemical Engineering Thermodynamics - I 3.76 P-V-T Behaviour and Heat Effects

Reactants are at 600 K.


∆H1 = ∑ HR = enthalpy of reactants at 600 K over 298 K
°
= ∑ ni Cpmi (600 – 298)
° °
= nCH4 · Cpm‚ CH (600 – 298) + nH2O · Cpm‚ H (600 – 298)
4 2O

= 1 × 44.026 (600 – 298) + 2 × 34.826 (600 – 298)


= 34330.76 J
∆H2 = ∑ Hp = enthalpy of products at 1300 K over 298 K
°
= ∑ ni Cpmi (1300 – 298)
° °
= [nCO · Cpm‚ CO + nCO2 · Cpm‚ CO2
° °
+ nH2O · Cpm‚ H2O + nH2 · Cpm‚ H2 ] [1300 – 298]
= [0.87 × 31.702 + 0.13 × 49.830 + 0.87 × 38.742 + 3.13 × 29.994] ×
[1002]
= 161968.7 J
For the calculation of ∆H°298 (total) , we have to consider both the reactions (1) and (2) and
the amounts of CH4 reacted by (1) and CO reacted by (2).
Moles of CH4 reacted by (1) = 1 mol
Moles of CO reacted by (2) = 1 – x = 1 – 0.87 = 0.13 mol
∑ ∆H° = (∆H° ) × amount of CH reacted + (∆H° )
298 298 1 4 298 2

× amount of CO reacted
(+205.813) (–41.166)
= 1 ×1+ 1 × 0.13
= (+205.813) + (–5.3516)
Total heat of
reaction at 298 K = 200.461 kJ = 200461 J
The net enthalpy change is
Q = ∆H = ∑ H + ∑ ∆H° – ∑ H
p 298 R

or ° – ∆H
Q = ∆H = ∆H2 + ∑ ∆H298 1

= 161968.7 + 200461 – 34330.76


= 328098.94 J ≈ 328099 J
Heat required for
the reactor = Q = 328099 J … Ans.
This much amount of the heat (enthalpy) should be supplied to the reactor in order to
maintain the temperature.
Chemical Engineering Thermodynamics - I 3.77 P-V-T Behaviour and Heat Effects

Example 3.23 :
The equation which connects U, P and V for several gases is
U = a + b PV
where a and b are constants. Prove that for a reversible adiabatic process, PVγ = constant,
where γ = (b + 1)/b.
Solution :
For a reversible adiabatic process, first law gives
0 = dU + PdV
The first law is dU = Q – W
Q = dU + W
Q = dU + PdV … (1)
for a reversible adiabatic process, Q = 0. Therefore, Equation (1) becomes
dU + PdV = 0
Rearranging gives,
dU
dV = – P … (2)
We have : U = a + bPV
Differentiating with respect to V yields
dU d(a + bPV) dP
dV = dV = bV dV + bP
dU  dP 
dV = b P + V dV … (3)
Equating (2) and (3), we get
–P = b P + V dV
dP
 
dP
bP + P + bV dV = 0
dP
P(b + 1) + bV dV = 0
Multiplying by dV and dividing by bPV, we get
b + 1 dV + dP = 0
 b  V P
b+1
Given : b = γ
dV dP
∴ γ V + P = 0
Integrating gives
γ ln V + ln P = constant
ln Vγ + ln P = constant
ln PVγ = constant
∴ PVγ = constant
Chemical Engineering Thermodynamics - I 3.78 P-V-T Behaviour and Heat Effects

Example 3.24 :
Using Boyle's law, Charle's law and taking differential of the relationship, V = f(P, T, n),
derive the ideal gas equation.
Solution :
V = f(P, T, n) … (1)
It means that the volume of a gas is a function of its pressure, temperature and number of
moles.
Volume is a state function, therefore, its total differential can be written as
∂V ∂V dT + ∂V dn
dV =   dP +    ∂n  … (2)
 ∂P T‚ n  ∂T P‚ n  T‚ P
Boyle's law at constant T and n :
PV = constant
PV = c
∴ V = c/P
Differentiating with respect to P at constant T and n gives
∂V P(0) – c(1) –c
 ∂P  = – = P2
 T‚ n P 2

PV –V
= – P2 = P since c = PV … (3)
Charle's law at constant P and n :
V
T = c
Differentiating with respect to temperature at constant P and n gives
V = cT
∂V = c
 ∂T 
 P‚ n
∂V = V
 ∂T  since c = V/T … (4)
 P‚ n T
Since at constant P and T, V ∝ n we can write
V = c'n
Differentiating with respect to n at constant P and T yields
∂V
 ∂n  = c'
 P‚ T
∂V V
 ∂n  = n since V = c'n … (5)
 P‚ T
Substituting for (∂V/∂P)T, n, (∂V/∂T)P, n and (∂V/∂n)P, T from Equations (3), (4) and (5)
respectively in Equation (2), it becomes
V dT dn
dV = – P dP + V T + V n
dV dP dT dn
∴ V = – P + T + n
Chemical Engineering Thermodynamics - I 3.79 P-V-T Behaviour and Heat Effects

Integrating and setting the constant of integration equal to ln R where R is the gas constant,
we get
ln V = –ln P + ln T + ln n + ln R
ln V + ln P = ln n + ln R + ln T
ln (PV) = ln (nRT)
∴ PV = nRT
which is the ideal gas equation.
Example 3.25 :
Calculate the molar volume and compressibility factor of isopropanol vapour at
473 K (200 °C) and 10 bar using the virial equation.
Data : The virial coefficients are :
B = –3.88 × 10–4 m3/mol and C = –2.6 × 10–8 m6/mol2
Solution :
P = 10 bar = 10 × 105 Pa
T = 473 K, R = 8.31451 m3·Pa/(mol·K)
B = –3.88 × 10–4 m3/mol and C = –2.6 × 10–8 m6/mol2
PV B C
We have : RT = 1 + V + V2
This equation can be rearranged for an iterative calculation as
Vi + 1 = 1 + V + 2 P
B C RT
 i V i
where i is the iteration number, Vi is the volume calculated in the ith iteration and Vi+1 is that in
the (i + 1)th iteration. To start the iterative calculation, when i = 0, Vo may be taken as equal to
the volume obtained with the help of the ideal gas law. Using Vo calculate V1. Compare Vo and
V1. If they are quite different, continue iterative calculations till the values of V are close enough.
Volume using the ideal gas law :
RT
Vo = P
8.3145 × 473
= = 3.9327 × 10–3 ≈ 3.933 × 10–3 m3/mol
10 × 105
First iteration : Vo = 3.933 × 10–3 m3/mol
V1 = 1 + V + 2  P
B C RT
 o V o
 3.88 × 10–4 2.6 × 10–8  8.3145 × 473
= 1 –
(3.933 × 10–3)2
– ×
 3.933 × 10 –3 10 × 105

= 1 –
3.88 × 10–4 2.6 × 10–8 
(3.933 × 10–3)2
– × 3.933 × 10–3
 3.933 × 10 –3

V1 = 3.5384 × 10–3 m3/mol


∴ Vcalculated ≠ Vassumed
Chemical Engineering Thermodynamics - I 3.80 P-V-T Behaviour and Heat Effects

Second iteration :
V1 = 3.5384 × 10–3 m3/mol

V1 = 1 + V + 2 P
B C RT
 1 V 1
 3.88 × 10–4 2.6 × 10–8  8.3145 × 473
= 1 – –  ×
 3.5384 × 10 –3 (3.5384 × 10 )
–3 10 × 105
= 3.4936 × 10–3 m3/mol
∴ V2 ≠ V1
Third iteration :
V2 = 3.4936 × 10–3 m3/mol

V3 = 1 + V + 2 P
B C RT
 2 V 2

= 1 –
3.88 × 10–4 2.6 × 10–8  8.31451 × 473
3.4936 × 10–3 (3.4936 × 10–3)2
– ×
  10 × 105
= 3.4878 × 10–3
Fourth iteration :
V3 = 3.4878 × 10–3 m3/mol

V4 = 1 + V + 2 P
B C RT
 3 V 3
 3.88 × 10–4 2.6 × 10–8  8.31451 × 473
= 1 – –  ×
 3.4878 × 10 –3 (2.487 × 10 ) 
–3 2 10 × 105
= 3.4871 × 10–3 m3/mol
As V4 ≈ V3, V4 is the final root and hence the correct volume.
The molar volume of isopropanol vapour is 3.4871 × 10–3 m3/mol. … Ans.
Compressibility V Vreal gas
factor = Z = (RT/P) = V
ideal gas
3.4871 × 10–3
= = 0.8866 … Ans.
3.933 × 10–3
Example 3.26 :
Determine the value of compressibility factor at the critical point (Zc) for a van der Waals
gas.
Solution :
The van der Waals equation of state is
P + a  (V – b) = RT
 V 2
RT a
P = V – b – V2
Chemical Engineering Thermodynamics - I 3.81 P-V-T Behaviour and Heat Effects

The critical isotherm shows a point of inflection at the critical point. At the point of
inflection,
 ∂P  = 0 and  ∂2P  = 0
∂V ∂V2
 T c
 T c
The van der Waals equation at the critical point is
RTc a
Pc = V – b – 2
c V c
Dropping the subscript c for simplification,
RT a
P = V – b – V2 … (1)
At the critical point T = Tc and is constant.
Differentiating Equation (1) with respect to V at constant T
 ∂P  = (V – b) × (0) – RT (1) – V2 × (0) – a (2V)
∂V  
 T (V – b)2 V4
 ∂P  = – RT + 2a = 0
∂V … (2)
 T (V – b)2 V3
 ∂2P  = – (V – b)2 (0) – RT [2(V – b)] + 2 V3 (0) – a (3V2)
∂V2    
 T (V – b)4 V6
 ∂2P  = 2RT – 6a = 0
∂V2 … (3)
 T (V – b)3 V4
–RT 2a
We have : (V – b)2 + V3 = 0
Multiplying this equation by 3/V gives
–3RT 6a
V(V – b)2 + V4 = 0
6a 3RT
∴ V4 = V(V – b)2 … (4)
2RT 6a
We have : (V – b)3 – V4 = 0 … (3)
6a
Replacing V4 from Equation (3) using Equation (4) gives,
2RT 3RT
(V – b)3 – V(V – b)2 = 0
2 3
V–b = V
2V = 3V – 3b
V = 3b
∴ b = V/3 … (5)
Chemical Engineering Thermodynamics - I 3.82 P-V-T Behaviour and Heat Effects

Substituting for b in Equation (2), we get


–RT 2a
(V – b)2 + V3 = 0
2a RT
V3 = (V – b)2
2a RT 9RT
V3 = (V – V/3)2 = 4V2
9RTV
∴ a = 8 … (6)
Substituting for a and b in Equation (1), we get
RT a
P = V – b – V2
RT 9RTV
P = V – V/3 – 8V2
3RT 9RT 12RT – 9RT
P = 2V – 8V = 8V
3RT
P = 8V
PV 3
∴ RT = 8
This is valid at the critical point.
PcVc 3
∴ RTc = 8 … (7)

The compressibility factor at the critical point, Zc is given by


Vc Volume of real gas
Zc = (RT /P ) = Volume predicted by ideal gas law
c c
PcVc
= RT … (8)
c
∴ From (7) and (8), we get
Vc 3
Zc = (RT /P ) = 8 … Ans.
c c
Example 3.27 :
Determine the van der Waals constants a and b, and the molar volume of ethane at the
critical point.
Data : Tc = critical temperature = 305.2 K
Pc = critical pressure = 49.4 bar
Solution :
Tc = 305.2 K and Pc = 49.4 bar = 49.4 × 105 Pa
Pc = 49.4 × 105 N/m2
Chemical Engineering Thermodynamics - I 3.83 P-V-T Behaviour and Heat Effects

The constants a and b are given as


2
27 R2 Tc
a = 64 P (N·m4) /mol2
c
RTc
and b = 8P m3/mol
c
R = 8.31451 m3·Pa/(mol·K)
= 8.31451 N·m/(mol·K)
27 (8.31451)2 (305.2)2
a =
64 (49.4 × 105)
a = 0.5499 ≈ 0.55 (N·m4)/mol4 … Ans.
8.31451 × 305.2
b =
8 × 49.4 × 105
b = 6.42 × 10–5 m3/mol … Ans.
Calculation of V (molar volume) at the critical point :
The ideal gas law is
PV = RT
RT
V = P
RTc
Vc = P Vc – volume at the critical point
c
8.31451 × 305.2
=
49.4 × 105
= 5.1368 × 10–4 m3/mol
The van der Waals equation of state is
P + a  (V – b) = RT … for 1 mole
 V 2
For iterative calculations, it may be rearranged as
RT
V–b = a
P + V2
RT
V = + b
P + a 
 V 2

RT
Vi+1 = +b
P + a2
 V
 i
where the subscript i denotes the iteration number.
Vi+1 is the volume calculated in the (i + 1)th iteration and Vi is that in the ith iteration.
For the first iteration, i = 0. Iteration starts with i = 0, Vi = Vo and Vo may be taken as
5.1368 × 10–4 m3/mol which is the volume obtained with the help of the ideal gas law. (In all
Chemical Engineering Thermodynamics - I 3.84 P-V-T Behaviour and Heat Effects

problems start iteration with V = Vo = volume calculated using the ideal gas law). Using Vo
calculate V1. Compare V1 and V0. If they are close enough, i.e., if Vi (calculated) = Vo (assumed)
then Vi is the final root and the calculations are terminated (the correct volume) and if, they are
quite different then calculate V2 and so on (and continue to convergence on the value).
Iteration-1 : i = 0, Vi = Vo = 5.1368 × 10–4 m3/mol
RT
V1 = +b
P + a 
 V 2
 0

where T = Tc = 305.2 K, P = Pc = 49.4 × 105 N/m2


a = 0.55 (N·m4)/mol2 and b = 6.42 × 10–5 m3/mol
8.31451 × 305.2
V1 = 0.55 + 6.42 × 10–5
49.4 × 10 +
5
(5.1368 × 10–4)2
= 4.254 × 10–4 m3/mol
∴ V1 ≠ V0
Iteration - 2 : i = 1, Vi = V1 = 4.254 × 10–4 m3/mol
RT
V2 = +b
P + a 
 V 2
 1

8.31451 × 305.2
= + 6.42 × 10–5
49.4 × 105 + 0.55/(4.254 × 10–4)2
= 3.82 × 10–4 m3/mol
V2 ≠ V1
Iteration-3 : Vi = V2 = 3.82 × 10–4 m3/mol
8.31451 × 305.2
V3 = + 6.42 × 10–5
49.4 × 105 + 0.55/(3.82 × 10–4)2
= 3.56 × 10–4 m3/mol
Iteration-4 : Vi = V3 = 3.56 × 10–4 m3/mol
V4 = 3.376 × 10–4
Iteration-5 : Vi = V4 = 3.376 × 10–4 m3/mol
V5 = 3.24 × 10–4 m3/mol
Iteration-6 : V6 = 2.396 × 10–4 ≈ 2.40 × 10–4 m3/mol
Iteration-7 : V7 = 2.389 × 10–4 m3/mol
Iteration-8 : V8 = 2.383 × 10–4
V8 ≈ V7
∴ V8 is the final root and value of V.
∴ Volume of ethane gas at the critical point
= 2.383 × 10–4 m3/mol … Ans.
Chemical Engineering Thermodynamics - I 3.85 P-V-T Behaviour and Heat Effects

Example 3.28 :
Express the Beattie-Bridgeman equation of state in the pressure explicit virial form.
The Beattie-Bridgeman equation of state is
c  bB0
PV2 = RT 1 – VT3 V + B0 – V – A0 1 – V
a
     
where A0, B0, a, b and c are constants.
Solution :
The pressure explicit form of the virial equation of state is
PV B C D
RT = 1 + V + V2 + V3 + … … (1)
The Beattie-Bridgeman equation of state is
c  bB0
PV2 = RT 1 – VT3 V + B0 – V – A0 1 – V
a
… (2)
     
where A0, B0, a, b and c are constants.
We have to express Equation (2) in the form of Equation (1) (i.e., in the pressure explicit
form).
Dividing each term of Equation (2) by RT, we get
PV2  c   bB0 A0  a
RT = 1 – VT3 V + B0 – V  – RT 1 – V
PV2 bB0 c cB0 bcB0 A0 aA0
RT = V + B0 – V – T 3 – VT 3 + VT 2 3 – +
RT RTV
Dividing each term of the above equation by V gives
PV B0 bB0 c cB0 bcB0 A0 aA0
RT = 1 + V – V2 – T3V – T3V2 + T3V3 – RTV + RTV2
Collecting the terms in 1/V, 1/V2 … on the RHS of the above equation yields
PV B0 c A0 aA0 bB0 cB0 bcB0
RT = 1 + – 3
V T V RTV RTV– + 2 – V 2 – TV3 2 + T 3V 3
PV  c A0  1 aA0 cB0 1 bcB0 1
RT = 1 + B0 – T3 – RT V +  RT – bB0 – T3  V2 +  T3  V3
… (3)
Equation (3) is the pressure explicit virial form of the Beattie-Bridgeman equation.
Example 3.29 :
Determine the second and third virial coefficients of the van der Waals equation. Take
a = RT.
Solution :
The virial equation of state as a power series in reciprocal volumes (the pressure explicit
form of the virial equation) is
PV B C D
RT = 1 + V + V2 + V3 + … … (1)
where B, C and D are known as the second, third and fourth virial coefficients, respectively and
V is the molar volume, m3/kmol.
Chemical Engineering Thermodynamics - I 3.86 P-V-T Behaviour and Heat Effects

The van der Waals equation of state for 1 mole of a gas is


P + a  (V – b) = RT … (2)
 V 2
a ab
PV – Pb + V – V2 = RT

a ab
PV = RT + Pb – V + V2 … (3)

Equation (2) can be written to give P as


RT a
P = V – b – V2 … (4)

To express Equation (3) in the pressure explicit form, as a power series in 1/V, substitute P
from Equation (4) into Equation (3).

PV = RT + V – b – V2 b – V + V2
RT a a ab
 
RTb ab a ab
PV = RT + V – b – V2 – V + V2

RTb a
PV = RT + V – b – V

Given : a = RT
RTb RT
PV = RT + V – b – V

Dividing both sides by RT,


PV b 1
RT = 1 + V – b – V … (5)

b
V – b can be rearranged as
b b b  b –1
V–b = = 1 –
V  V
V 1 – V
b
 
With this Equation (5) becomes
PV b  b –1 1
RT = 1 + V 1 – V – V … (6)

Expand 1 – V : Based on the Binomial theorem (1 – x)–1 can be expanded (as a power
b –1
 
series in x) as
(1 – x)–1 = 1 + x + x2 + x3 + …
Chemical Engineering Thermodynamics - I 3.87 P-V-T Behaviour and Heat Effects

b
with x = V

1 – b –1 = 1 + b + b + …
2
∴ … (7)
 V V V2

Replacing 1 – V from the RHS of Equation (6) using Equation (7), we get
b –1
 
= 1 + V 1 + V + V2 + … – V
PV b b b2 1
RT  
PV 1 b b2 b3
RT = 1 – V + V + V2 + V3 + …
PV (b – 1) b2 b3
RT = 1 + V + V2 + V3 + … … (8)
PV B C D
We have : RT = 1 + V + V2 + V3 … (1)
Comparing Equations (8) and (1) term by term (i.e., comparing the coefficients of 1/V, 1/V2,
1/V3 terms of Equations (8) and (1), we get
For 1/V : b–1 = B  B = b–1 … Ans.
2
For 1/V : 2
b = C  C = b 2 … Ans.
3
For 1/V : 3
b = D  D = b 3 … Ans.
[Binomial theorem for any index :
n(n – 1) 2 n(n – 1) (n – 2) 3
(1 + x)n = 1 + nx + 2! x + 3! x + ……
n(n – 1) 2 n(n – 1) (n – 2) 3
(1 – x)n = 1 – nx + 2! x – 3! x +…

Based on these,
∴ (1 – x)–1 = 1 + x + x2 + x3 + x4 + …
∴ (1 + x)–1 = 1 – x + x2 – x3 + x4 – …]
Example 3.30 :
Show that the virial coefficients B', C' and D' of the volume explicit virial equation (Berlin
form) are related to B, C and D of the pressure explicit virial equation (Leiden form) by the
following relations :
B C – B2 D – 3BC + 2B3
B' = RT , C' = (RT)2 and D' = (RT)3
Solution :
The virial equation of state expressed as a power series in reciprocal volumes, i.e., the
pressure explicit form/Leiden form of the virial equation is
PV B C D
RT = 1 + V + V2 + V3 + … … (1)
where B, C and D are known as the second, third and fourth virial coefficients and V is the molar
volume, m3/kmol.
Chemical Engineering Thermodynamics - I 3.88 P-V-T Behaviour and Heat Effects

The volume explicit form/Berlin form of the virial equation of state is


PV
RT = 1 + B'P + C'P + D'P + …
2 3 … (2)
where B', C' and D' are known as the second, third and fourth virial coefficients.
We have obtain the relations between these two sets of coefficients as :
B C – B2 D – 3BC + 2B3
B' = RT , C' = (RT)2 and D' = (RT)3
These relations are obtained by elimination of P on the RHS of Equation (2) through use of
Equation (1) and comparing the resulting equation (which is a power series in 1/V) term by term
with Equation (1).
PV B C D
RT = 1 + V + V2 + V3 … (1)
Rearranging gives
P = 1 + V + V2 + V3 V
B C D RT
… (3)
 
Elimination of P from the RHS of Equation (2) :
Substituting P from Equation (3) into Equation (2), we get
1 + B + C + D  RT + C' 1 + B + C + D 2 (RT)
PV 2
= 1 + B '
RT  V V V V 2 3
 V V V V
2 3 2

+ D' 1 + V + V2 + V3 V3
B C D 3 (RT)3
 
PV B'RT B'B RT B'C RT B'D RT
RT = 1 + V + V2 + V3 + V4

+ C' 1 + V2 + V4 + V6 + V + V2 + V3 + V3 + V4 + V5  V2
B C D 2B 2C 2D 2BC 2BD 2CD (RT)2
2 2 2

 
← expanded square term →
+ D' 1 + V + V2 + V3 1 + V + V2 + V3 V3
B C D 2 B C D (RT)3
   
← cubic term is rewritten as →
PV B'RT B'B RT B'C RT B'D RT
RT = 1 + V + V 2 + V 3 + V4

+ C' 1 + V2 + V4 + V6 + V + V2 + V3 + V3 + V4 + V5  V2
B2 C2 D2 2B 2C 2D 2BC 2BD 2CD (RT)2
 

+ D' 1 + V2 + V4 + V6 + V + V2 + V3 + V3 + V4 + V5  ×
B2 C2 D2 2B 2C 2D 2BC 2BD 2CD
 

× 1 + V + V2 + V3 V3 … (4)
B C D (RT)3
 
Neglect the terms from 1/V4 including 1/V4.
Take only first terms from two brackets associated with D'.
Chemical Engineering Thermodynamics - I 3.89 P-V-T Behaviour and Heat Effects

For comparing Equation (4) with Equation (1) - both are power series in 1/V (reciprocal
volumes) – term by term we take from Equation (4) only the terms upto 1/V3 (i.e., 1, 1/V, 1/V2
and 1/V3 terms). These terms are underlined in Equation (4).
PV B'RT B'BRT B'CRT C'(RT)2 2C'B(RT)2 D'(RT)3
RT = 1 + V + V 2 + V 3 + V 2 + V 3 + V3 … (5)

We have :
PV B C D
RT = 1 + V + V2 + V3 … (1)

Comparing Equations (5) and (1) term by term gives


[Comparing coefficients associated with 1/V, 1/V2 and 1/V3 terms on the RHS of
Equations (5) and (1)]
1/V : B'RT = B
B
∴ B' = RT … Ans.

1/V2 : B'BRT + C'(RT)2 = C


 B  BRT + C'(RT)2 = C since B' = B/RT
RT
B2 + C'(RT)2 = C
C – B2
∴ C' = (RT)2 … Ans.

1/V3 : B'CRT + 2C'B (RT)2 + D'(RT)3 = D


We have : B' = B/RT and C' = (C – B2)/(RT)2
B 2(C – B2) B
∴ RT CRT + (RT)2 + D'(RT)3 = D
BC + 2BC – 2B3 + D'(RT)3 = D
3BC – 2B3 + D'(RT)3 = D
D'(RT)3 = D – 3BC + 2B3
D – 3BC + 2B3
∴ D' = (RT)3 … Ans.
Example 3.31 :
The Dieterici equation of state for 1 mole of a gas is given by
P(V – b) exp (a/RT V) = RT
where a and b are constants. Obtain the relations to find a and b in terms of critical
temperature (Tc) and critical pressure (Pc).
Solution :
The Dieterici equation of state is
P(V – b) ea/RTV = RT … (1)
Chemical Engineering Thermodynamics - I 3.90 P-V-T Behaviour and Heat Effects

This can be rearranged as


RT
P = (V – b) ea/RTV

Taking the natural logarithm of both sides gives

ln P = ln (V – b) ea/RTV = ln RT – ln [(V – b) · ea/RTV]


RT
 
ln P = ln RT – ln (V – b) – ln ea/RTV
a
ln P = ln RT – ln (V – b) – RTV … (2)

We know that at the critical point, the critical isotherm must show a point of inflection.
At the point of inflection :
 ∂P  = 0 and  ∂2P  = 0
∂V ∂V2
 Tc  Tc
Dropping the subscript c for the time being, we have
 ∂P  = 0 and  ∂2P  = 0
∂V ∂V2
 T  T
Differentiating Equation (2) w.r.t. V keeping T constant gives
a
ln P = ln RT – ln (V – b) – RTV

1  ∂P  = 0 – 1 × 1 – a – 1 
∂V RT  V2
P  T V–b

1  ∂P  = – 1 + a 
∂V … (3)
P   T  V – b RTV2

 ∂P  = P – 1 + a 
∂V  V – b RTV2 … (4)
 T

Differentiating Equation (3) w.r.t. V at constant T gives

1  ∂P  = – 1 + a  … Equation (3)
∂V  V – b RTV2
P  T

1  ∂2P   ∂P   1  ∂P    –1  a  1 
P ∂V2T + ∂VT – P2 ∂VT  = – (V – b)2 + RT – V4 × 2V

1  ∂2P  – 1  ∂P  2 = 1 2a
∂V2     –
P  T P ∂VT (V – b) 2 RTV3
Chemical Engineering Thermodynamics - I 3.91 P-V-T Behaviour and Heat Effects

1  ∂P 
Substituting for P ·   from Equation (3), the above equation becomes
∂VT
1  ∂2P  – 1 + a 2 = 1 2a

P ∂V2T –
 V – b RTV2 (V – b)2 – RTV3

1  ∂2P  – 1 + a 2 + 1 2a

P ∂V T
2  =
 V – b RTV2 (V – b)2 – RTV3

 ∂2P 
∂V2 = P (V – b)2 – (V – b) RTV2 + R2 T2 V4 + (V – b)2 – RTV3
1 2a a2 1 2a

 T
 ∂2P  = P  2 2a a2 2a 
∂V2 
– + 2 V4 – RTV3

 T (V – b) 2 (V – b) RTV 2 R2 T
At the critical point, Equations (4) and (5) are equal to zero. Therefore, writing Vc and Tc to
represent the molar critical volume and critical temperature, respectively, it follows that
1 a
– V –b + 2 = 0 … from (3) … (6)
c RT V c c
2 2a a2 2a
and (Vc – b) 2 – 2 + 2 4 – 3 = 0 … since P ≠ 0
(Vc – b) RTcVc R2 Tc Vc RTcVc
… from (4) … (7)
From Equation (6) :
1 a
Vc – b = 2
RTc Vc
2
RTcVc
∴ (Vc – b) = a … (8)

Substituting for (Vc – b) from Equation (8), Equation (7) becomes


2a2 2a2 a2 2a
2 4 – 2 2 + 2 4 – 3 = 0
R2 T c Vc R2 T c Vc R2 T c Vc RTcVc
a2 2a
∴ 2 4 = 3
RTcVc RTcVc

a
RTcVc = 2
a
∴ Vc = 2RT … (9)
c

At the critical point, the Dieterici equation of state is given by


a/RTcVc
Pc(Vc – b) e = RTc … (10)
Chemical Engineering Thermodynamics - I 3.92 P-V-T Behaviour and Heat Effects

Substituting Vc from Equation (9) into Equation (10), we get

Pc 2RT – b · e c
a a/RT × (a/2RTc)
= RTc
 c 

Pc 2RT –b e2 = RTc


a
 c 
a RTc
2RTc – b = Pc · e2
a RTc
∴ b = 2RT – P e2
c c

We have : e2 = 7.389
a RTc
∴ b = 2RT – 7.389 P … (11)
c c
1 a
We have : Vc – b = RT V2
c c

Substituting for b this becomes


1 a
RTc = 2
a RTcVc
Vc – 2RT + 7.389 P
c c

2  a RTc 
RTc Vc = a Vc – 2RT + 7.389 P 
 c c

a
We know that Vc = 2RT . With this the above equation becomes
c

RTc · a2 a2 aRTc
2 = aVc – 2RT + 7.389 P
4R2 Tc c c

a2 a· a a2 aRTc
4RTc = 2RTc – 2RTc + 7.389 Pc

a2 a2 a2 aRTc
4RTc = 2RTc 2RTc + 7.389 Pc

a2 aRTc
∴ 4RTc = 7.389 Pc
2
4R2 Tc
a = 7.389 P
c
2
0.54134 R2 Tc
a = Pc … Ans.
Chemical Engineering Thermodynamics - I 3.93 P-V-T Behaviour and Heat Effects

Substituting for a, Equation (11) becomes


a RTc
b = 2RT – 7.389 P
c c
2
0.54134 R2 Tc RTc
b = 2RT · P – 7.389 P
c c c
0.27067 RTc 0.13534 RTc
b = Pc – Pc
0.13533 RTc
b = Pc … Ans.

Example 3.32 :
The Berthelot equation of state for 1 mole of a gas is given by
P + a  (V – b) = RT
 TV2
where a and b are constants, characteristic of the gas. Obtain relations to evaluate the constants
a and b in terms of Tc and Pc.
Solution :
The Berthelot equation of state is
P + a  (V – b) = RT … for 1 mole of gas
 TV2
This equation can be written in the form
RT a
P = V – b – TV2 … (1)

We know that at the critical point, the critical isotherm must show a point of inflection.
At the point of inflection,
 ∂P  = 0 and  ∂2P  = 0
∂V ∂V2
 Tc  Tc
Dropping the subscript c, we have
 ∂P  = 0 and  ∂2P  = 0
∂V ∂V2
 T  T
Differentiating Equation (1) w.r.t. V at constant T (i.e., Tc) gives

 ∂P  = (V – b) × 0 – RT × 1 – a –2V
∂V T  V4 
 T (V – b)2

 ∂P  = –RT + 2a
∂V … (2)
 T (V – b)2 TV3
Chemical Engineering Thermodynamics - I 3.94 P-V-T Behaviour and Heat Effects

Differentiating Equation (2) w.r.t. V at constant T gives


 ∂2P  = (V – b)2 × 0 – (–RT) × 2 (V – b) × 1 +2a V3 × 0 – 1 × 3V2
∂V2 T  
 T (V – b)4 V6
 ∂2P  = 2RT – 6a
∴ ∂V2 … (3)
 T (V – b)3 TV4
At the critical point, Equations (2) and (3) are equal to zero. Therefore, writing Vc and Tc to
represent the molar critical volume and critical temperature, respectively, it follows that
–RTc 2a
(Vc – b) 2 + 3 = 0 … (4)
T V c c
2RTc 6a
and (Vc – b)3 – T V4 = 0 … (5)
c c

At the critical point, the Berthelot equation is given by


P + a  (V – b) = RT
 c T V2  c c … (6)
 c c

From Equation (4) we have


RTc 2a
(Vc – b)2 = T V3 … (7)
c c
Dividing both sides by (Vc – b) gives
RTc 2a
(Vc – b)3 = T V3 (V – b) … (8)
c c

From Equation (5), we have


2RTc 6a
(Vc – b)3 = T V4
c c
RTc 3a
(Vc – b)3 = T V4 … (9)
c c

From Equations (8) and (9) we have (equate the right hand sides of these equations as the left
hand sides of these equations are the same).
2a 3a
3 = 4
Tc Vc (Vc – b) Tc Vc
2 3
(Vc – b) = Vc
2Vc = 3 (Vc – b) = 3Vc – 3b
∴ Vc = 3 b … (10)
Chemical Engineering Thermodynamics - I 3.95 P-V-T Behaviour and Heat Effects

Substituting this value of Vc in Equation (7), we get


RTc 2a
(3b – b) 2 = Tc (3b)3
RTc 2a R Tc 2a
4b2 = 27b3 Tc , 4 = 27bTc
2 8a
∴ Tc = 27bR … (11)
We have :
P + a  (V – b) = RT
 c T V2  c c
 c c
RTc a
Pc = V – b – 2
c Tc Vc
Substituting Vc = 3b, we get
RTc a
Pc = 3b – b – 9b2 T
c
RTc a
Pc = 2 b – 9b2 T
c
2
9bTc (RTc) – 2a 9bR Tc – 2a
Pc = 2
18b Tc = 18 b2 Tc
2
Substituting Tc from Equation (11) the above equation gives

9bR 27bR – 2a
8a 8a
  3 – 2a
Pc = 18b2 Tc = 18 b2 Tc
8a – 6a 2a
Pc = 54 b2 T = 54 b2 T … (12)
c c
From Equation (11), we have
8a
b = 2
27 R Tc
Substituting this value of b in Equation (12), we get
2a
Pc =
54 Tc  8a 2
2
27 R Tc
2 4
2a (27R Tc)2 2 × 729 R2 Tc
Pc = =
54 Tc × 64a2 54 × Tc × 64a
3
27 R2 Tc
= 64a
3
27 R2 Tc
∴ a = 64 P … Ans.
c
Chemical Engineering Thermodynamics - I 3.96 P-V-T Behaviour and Heat Effects

8a
We have : b = 2
27 R Tc
Substituting for a it becomes
3
8 × 27 R2 Tc 1 RTc
b = 64 Pc × 2 =
8Pc … Ans.
27 R Tc
Example 3.33 :
Show that at moderate and low pressures, the van der Waals equation can be written in the
form as given below :
PV = RT (1 – BP)

B = RT RT – b
1 a
where,
 
Solution :
The van der Waals equation of state is
P + a  (V – b) = RT
 V 2
a ab
PV – Pb + V – V2 = RT
a ab
PV = RT + Pb – V + V2
Neglecting the last term ab/V2 (which is small if the pressure is not large - it consists of the
product of two small quantities), we get
a
PV = RT + Pb – V

Replacing V in a/V as an approximation, by the ideal gas value RT/P, the above equation
becomes
aP
PV = RT + Pb – RT

Taking RT common from the terms on RHS gives

PV = RT 1 + RT – R2 T2
Pb aP
 

= RT 1 – P R2 T2 – RT
a b
  

= RT 1 – RT RT – b = RT 1 – P × RT RT – b


P a 1 a
     

B = RT RT – b . Therefore,
1 a
Given :
 
PV = RT [1 – BP] … Ans.
ËËË
Chapter ... 4
SECOND LAW OF THERMODYNAMICS
The energy balance of a particular process is based upon the first law of thermodynamics.
The total energy entering a process plus any energy addition during the process is equal to the
total energy leaving the process. The first law of thermodynamics thus helps us to determine the
energy changes involved in a process. According to the first law of thermodynamics, all forms of
energy are equivalent and all processes are possible provided that energy can neither be created
nor destroyed. But this is not the case-all forms of energy are not equivalent qualitywise (heat is
a low grade energy and electrical energy is a high grade energy) and all processes are not
practically possible (A hot cup of tea cools because of transfer of heat to the cooler surroundings.
According to the first law, heat lost by the tea is equal to heat gained by the surroundings. This
process of cooling of tea is possible. If the reverse process to occur, the tea should get heated by
absorbing heat from the cooler surroundings. In this process also the first law will be equally
satisfied, but such a process does not occur in practice). Therefore, the first law of
thermodynamics, though exact, has certain limitations. The first law of thermodynamics does not
provide any information regarding the feasibility of a process, the conditions under which the
process occurs, the extent of process and the quality of energy.
It is the second law of thermodynamics which helps us (i) to predict whether a particular
process or chemical reaction is feasible or not under the specified conditions, (ii) to access how
far a process or chemical reaction proceeds, i.e., to assess the extent of a process or chemical
reaction and (iii) to find the effect of operating conditions (P, T) on the extent of a process or
chemical reaction.
The second law of thermodynamics has been expressed in many ways. All the statements
given below are more or less indicative of one and the same meaning (i.e., they all are
equivalent).
1. Heat or in general any type of energy flows from a higher level to a lower level.
2. When two bodies are at different temperatures, heat flows from a hot body to a
relatively cold body.
3. The energy of the universe is constant but the entropy of the universe is continuously
increasing.
4. The Clausius statement of the second law of thermodynamics states that it is impossible
to construct a device, operating in a cycle, which will produce no effect other than the
transfer of heat from a low temperature body to a high temperature body. It means that
energy in the form of work must be supplied to a device for transfer of heat from a cold
body to a hot body.
OR : It is impossible to transfer heat from a cold body to a hot body without the aid of
external work.
(4.1)
Chemical Engineering Thermodynamics - I 4.2 Second Law of Thermodynamics

OR : Heat cannot of its own accord flow from a body at a lower temperature to a body at a
higher temperature. Some work must be expended for doing this.
5. In a reversible process the entropy remains constant but in an irreversible process the
entropy increases.
6. All natural or spontaneous processes are not thermodynamically reversible.
7. All natural or spontaneous processes can be made to do work but the maximum work
can only result from a reversible process.
8. All processes occurring in nature are associated with a gain of entropy of the system and
its surroundings.
9. Complete conversion of heat into work is impossible without producing some changes
in the system or its surroundings.
10. As a result of natural processes, energy steadily becomes less available for the
performance of work.
11. The Kelvin-Planck statement of the second law of thermodynamics states that it is
impossible to construct a device, operating in a cycle, which will produce no effect
other than the extraction/absorption of heat from a single thermal reservoir and the
performance of an equivalent amount of work.
OR : It is impossible to take heat from a hot reservoir and convert it completely into work,
without losing some of it to a cold reservoir.
The Kelvin-Planck statement thus implies that no heat engine is 100% efficient.
Out of the statements given above, the Kelvin-Planck and Clausius statements are widely
used.
A perpetual motion machine of the second kind, abbreviated as PMM2, is a cyclically
operating device which absorbs heat from a single thermal reservoir and delivers an equivalent
amount of work. The Kelvin-Planck statement of the second law of thermodynamics tells us that
a PMM2 is impossible.
Like the first law of thermodynamics, the second law of thermodynamics rests on
experimental evidence.
Heat Engine :
• A heat engine is a cyclically operated device for the conversion of heat into work.
• A heat engine is a device or machine which produces work from heat in a cyclic
process.
• A heat is an energy conversion device which converts heat (a low grade energy) into
work (a high grade energy). For example, a steam power plant.
Working Medium :
A fluid which absorbs and rejects heat in a heat engine and undergoes a cyclic change is
called a working medium.
The working medium in a steam power plant is steam.
Thermal reservoir/Heat reservoir :
A body which is capable of absorbing or rejecting an infinite quantity of heat without change
in its temperature is called a thermal reservoir or a heat reservoir.
Chemical Engineering Thermodynamics - I 4.3 Second Law of Thermodynamics

It is a storage or reservoir of infinite heat energy content whose temperature does not change
with the flow of heat in or out.
Source and Sink :
• A thermal reservoir from which the working medium of a heat engine absorbs heat is
called a source or a hot reservoir.
• It is a thermal reservoir at higher temperature from which heat is extracted.
• A high temperature reservoir from which heat is absorbed by the system is called a
source.
• A typical source is a constant temperature furnace where fuel is continuously burnt.
• A thermal reservoir to which the working medium of a heat engine rejects heat is called
a sink or a cold reservoir.
• It is a thermal reservoir at lower temperature to which heat is rejected.
• A low temperature thermal reservoir to which heat is rejected by the system is called a
sink.
• A typical sink is water from a river or lake or air from the atmosphere.
The primary objective of a heat engine is to deliver net work by absorbing energy in the form
of heat. A typical example of a heat engine is a steam power plant shown in Fig. 4.1. The
components of this plant are : boiler, turbine, condenser and pump. The working fluid, steam,
passes in a cycle through the system from the boiler through the turbine to the condenser and
back to the boiler through the pump.
Superheated
steam

WT
Boiler Turbine
Q1
from furnace
Q2

Condenser

Saturated liquid
Pump water

WP

Fig. 4.1 : Steam power plant as a heat engine (flow diagram)


Superheated steam is generated from saturated liquid water in the boiler (by receiving heat
from the furnace) and is used to produce work in the steam turbine. The low pressure steam
leaving the turbine is condensed in the condenser by rejecting heat to the cooling water and the
saturated liquid water leaving the condenser is fed back to the boiler with the help of pump to
complete the cycle.
Chemical Engineering Thermodynamics - I 4.4 Second Law of Thermodynamics

WT is the work output of the turbine and WP is the work supplied to the pump.
Q1 is the heat supplied to the boiler from the furnace and Q2 is the heat rejected by the steam
in the condenser to the cooling water.
The net heat interaction between the system and its surroundings for a completed cycle is
Qnet = Qboiler + Qcondenser
= Q1 + (– Q2) = Q1 – Q2 , by the sign convention … (4.1)
The quantity Qnet is the algebraic sum of the energy added to the system as high temperature
heat and the energy withdrawn from the system as low temperature heat.
Similarly, the net work interaction between the system and its surroundings, for a completed
cycle is
Wnet = W = Wturbine + Wpump
= WT – WP, by the sign convention … (4.2)
For a cyclic process, according to the first law,

O dQ = ⌠
⌡ ⌡
O dW

or [∑ Q]cycle = [∑ W]cycle , [since (∑ ∆E)cycle = 0]


Qnet = Wnet (= W)
i.e., Q1 – Q2 = WT – WP … (4.3)
Hence forth, we will use the symbol W for Wnet, i.e., for a net amount of work delivered or
produced.
A heat engine is an energy conversion device/machine. Therefore, its performance is
expressed in terms of thermal efficiency or simply efficiency (η).
The efficiency of a heat engine is defined as the ratio of the net work output/net work
delivered to the heat input/heat added/heat absorbed from the source. That is,
W
η = Q … (4.4)
1

Thus, the efficiency of a heat engine is the fraction of the heat added/heat absorbed which is
converted to net work output/delivered.
Substituting for W from Equation (4.3) in Equation (4.4), we get
Q1 – Q2 Q2
η = Q1 = 1 – Q1 … (4.5)

From Equation (4.5) it is clear that for η to be unity (100 percent efficiency), Q2 must be
zero. Such a engine is not possible as some heat has to be rejected to the sink. Therefore, the
efficiency of a heat engine is always less than unity.
For thermodynamic analysis, a heat engine can be schematically represented as shown in
Fig. 4.2, wherein the heat engine exchanging heat with a source and a sink and delivering a net
amount of work W in a cycle is represented by a circle and the source and sink are represented by
rectangular blocks.
Chemical Engineering Thermodynamics - I 4.5 Second Law of Thermodynamics

High temperature
thermal reservoir
(source) at T1
High temperature thermal reservoir
Source Hot reservoir
Q1

Heat
W
engine

Q2

Low temperature Low temperature thermal reservoir


thermal reservoir Sink Cold reservoir
(sink) at T2

Fig. 4.2 : Schematic representation of a heat engine


Refrigerator and Heat pump :
Heat engines are work producing devices, whereas refrigerators and heat pumps are work
consuming devices. Though both refrigerator and heat pump consume energy in the form of work
(to reject heat to a higher temperature), their objectives are different - the objective of a
refrigerator is to maintain a body at a temperature lower than the temperature of the ambient
atmosphere, while the objective of a heat pump is to maintain a body at a temperature higher than
the temperature of the ambient atmosphere.
A device which, operating in a cycle, absorbs heat from a body at low temperature and
rejects heat to a body at high temperature when work is done on it from an external source (i.e.,
with the help of work input) is called a heat pump or a refrigerator.
Refrigerator :
A device which, operating in a cycle, maintains a body at low temperature by absorbing heat
from the body and rejecting heat to the ambient atmosphere when work is done on it is called a
refrigerator.
A refrigerator is a device which, operating in a cycle, maintains a body at a temperature
lower than the temperature of the ambient atmosphere (by transferring heat from the body to the
ambient atmosphere with the help of work input that increases the temperature potential of the
heat to be transferred).
Consider a refrigerator [Fig. 4.3 (a)] which maintains a body, say A, at a temperature T2
lower than the ambient atmosphere temperature T1. Even though the body A is well insulated,
heat Q2 naturally flows or leaks into the body A from the ambient atmosphere since it is at a
lower temperature than the ambient atmospheric temperature (T2 < T1). In order to maintain the
body A at constant temperature T2, heat Q2 has to be removed from it at the same rate at which
heat flows or leaks into the body A. In the refrigerator, heat Q2 is absorbed by a working fluid
Chemical Engineering Thermodynamics - I 4.6 Second Law of Thermodynamics

called the refrigerant. The refrigerant evaporates in an evaporator at a temperature lower than T2
by absorbing the latent heat of vaporisation from the body A, which is cooled or refrigerated. The
vapour from the evaporator is compressed by a compressor by consuming work W and is then
condensed in a condenser by rejecting the latent heat of condensation of the refrigerant Q1 at a
temperature higher than the ambient atmospheric temperature T1. The liquid refrigerant then
passes through an expansion valve in which a Joule-Thomson expansion occurs, resulting in the
evaporation of some liquid and cooling of both liquid and vapour formed to the temperature of
the evaporator which is less than T2.
The refrigerant enters the evaporator where the rest of the liquid is converted to vapour by
absorption of heat Q2 from the body A (since the temperature of refrigerant is less than T2).
Atmosphere at T1

Q1 Atmosphere
T1

Condenser Q1

Expansion Refrigerator W
valve
Compressor W

Q2
Evaporator

Body A at T2
Q2

Body A at T2

Q2 (leakage from atmosphere)

(a) Flow diagram (vapour compression) (b) Schematic representation


Fig. 4.3 : Refrigerator/Refrigeration unit
In case of a refrigerator, the purpose is to maintain the body A at low temperature T2 by
removing heat Q2 from it by consuming work W and therefore, Q2 and W are of primary interest.
The performance of a refrigerator is expressed in terms of coefficient of performance (COP).
The coefficient of performance of a refrigerator is defined as the ratio of the desired
refrigerating effect to the work input or supplied to achieve the effect.
Refrigerating effect Heat removed/absorbed from the cold body
(COP)R = Work input/supplied = Work input
Q2
(COP)R = W … (4.6)
Chemical Engineering Thermodynamics - I 4.7 Second Law of Thermodynamics

According to the first law for a refrigerator,


Q2 + W = Q1
∴ W = Q1 – Q2
Q2
(COP)R = Q – Q … (4.7)
1 2

Heat pump :
A device which, operating in a cycle, maintains a body at a temperature higher than the
ambient atmosphere by transferring heat from the ambient atmosphere to the body when work is
done on it by an external source is called a heat pump.
A heat pump is a device which, operating in a cycle, pumps heat from a low to a high
temperature with the help of work input.
A refrigerating unit, used for cooling (air conditioning) during the warm weather, can be
used as a heat pump for heating (homes) during the cold season.
Suppose that a body B is to be maintained at a temperature T1 higher than the
temperature, T2, of the ambient atmosphere/surroundings. Though the body B is well insulated,
heat Q1 will flow out from the body to the surroundings because of the temperature difference
(T1 – T2, T1 > T2) and in order to maintain body B at T1 we have to supply heat Q1 to it. For this,
work W is supplied to a heat pump which removes heat Q2 from the ambient atmosphere and
supplies heat Q1 to the body B.

Body B Q1
at T1

Q1

W + Q2 = Q1
W Heat
pump W = Q1 - Q2

Q2

Atmosphere
at T2

Fig. 4.4 : A heat pump


As the purpose of a heat pump is to maintain the body at a temperature T1 higher than the
temperature of the surroundings by supplying heat Q1 to the body, Q1 and W are of primary
interest.
Chemical Engineering Thermodynamics - I 4.8 Second Law of Thermodynamics

Therefore, the coefficient of performance of a heat pump is given by


Heat supplied to the body at high temperature
(COP)HP = Work input (supplied) to supply heat
Q1
= W … (4.8)

According to the first law, for the heat pump as a system,


Q2 + W = Q1
W = Q1 – Q2
Q1
(COP)HP = Q – Q … (4.9)
1 2

Subtracting Equation (4.7) from Equation (4.9), we get


Q1 Q2 Q1 – Q2
(COP)HP – (COP)R = Q – Q – Q – Q = Q – Q = 1
1 2 1 2 1 2

(COP)HP = (COP)R + 1 … (4.10)


That is, the COP of a heat pump is greater than the COP of a refrigerator by unity.
For thermodynamic analysis, we can represent a heat engine and a heat pump or refrigerator
as shown in Fig. 4.5.

High temperature High temperature


thermal reservoir thermal reservoir
(source) at T1 at T2

Q1 Q1

Heat Heat pump


W W or
engine
Refrigerator

Q2 Q2

Low temperature Low temperature


thermal reservoir reservoir
[ sink ] at T2 at T2

(a) Schematic representation of a (b) Schematic representation of a


heat engine heat pump/refrigerator
Fig. 4.5
• Perpetual motion machine of the second kind [PMM2] :
A cyclically operating device which absorbs heat continuously from a single thermal
reservoir and converts it completely into work is called a pertual motion machine of the second
kind.
Chemical Engineering Thermodynamics - I 4.9 Second Law of Thermodynamics

Thermal reservoir

Q1

Heat
W
engine

Q2 = 0
Fig. 4.6 : A PMM2
For a heat engine, we have
W
W = Q1 – Q2 and η = Q
1

If Q2 = 0, W = Q1 and therefore, η = 1. That is, if a heat engine is operating with only one
reservoir, it is 100% efficient. This is a violation of the Kelvin-Planck statement since according
to the Kelvin-Planck statement no engine is 100% efficient. Such a engine is called a PMM2 and
is thus impossible.
A heat engine operates, therefore, with a minimum of two thermal reservoirs, absorbing heat
from a high temperature thermal reservoir or a source, converting a part of it into work and
rejecting the remaining part of it to a low temperature reservoir or a sink.
• The impossible system and possible system as per the Kelvin-Planck statement of the second
law of thermodynamics are shown in Fig. 4.7 (a) and (b), respectively.
High temperature
High temperature
thermal reservoir
thermal reservoir
(source)
(source) at T1
at T1

Q1 Q1

Heat Heat
W W
engine engine

Q2

Low temperature
thermal reservoir
[ sink ] at T2

(a) Impossible heat engine as per (b) Possible heat engine as per
Kelvin-Planck statement Kelvin-Planck statement
Fig. 4.7
Chemical Engineering Thermodynamics - I 4.10 Second Law of Thermodynamics

• The impossible system and possible system as per the Clausius statement of the second law
of thermodynamics are shown in Fig. 4.8 (a) and (b) respectively.
Source at T1 Source at T1

Q2 Q1 Q = Q + W
1 2

Heat Heat
W
pump pump

Q2 T1 > T2 Q2

Sink at T2 Sink at T2

(a) Impossible system (Heat pump) as per (b) Possible system (Heat pump) as per
Clausius statement Clausius statement
Fig. 4.8
Consider an arrangement as shown in Fig. 4.8 (a) consisting of a heat pump, a source and a
sink. The heat pump shown in Fig. 4.8 (a) transfers heat Q2 from the sink to the source without
absorbing any work. This is impossible as per the Clausius statement of second law of
thermodynamics. Consider an arrangement as shown in Fig. 4.8 (b) which comprises of a heat
pump, a source and a sink. The heat pump transfers heat Q2 from the sink and supplies
heat Q2 + W to the source with work input of W. This is not contradictory to the Clausius
statement. Therefore, it is a possible system.
Equivalence of Kelvin-Planck and Clausius Statements :
Although the Clausius and Kelvin-Planck statements of the second law of thermodynamics
appear to be different, they, in fact, are equivalent.
The equivalence of the Clausius and Kelvin-Planck statements can be proved by showing
that violation of either of the statements leads to the violation of the other.
(1) Violation of Kelvin-Planck statement results into violation of Clausius statement :
Consider a heat engine as shown in Fig. 4.9 (a). It absorbs heat Q1 from a source and converts it
entirely into work W. According to the Kelvin-Planck statement such a heat engine is impossible
and therefore, it violates the Kelvin-Planck statement.
Let us assume a heat pump which, operating between the same two reservoirs (source and
sink), removes heat Q2 from the sink and delivers heat Q1 + Q2 (i.e., Q2 + W) to the source by
receiving work W (W = Q1) from the engine. Hence, we see that the combined system of heat
engine and heat pump, i.e., the heat engine and heat pump together acts as a heat pump. Fig. 4.9
(b) which, operating in a cycle, produces the sole effect of removing heat from a sink and
Chemical Engineering Thermodynamics - I 4.11 Second Law of Thermodynamics

transferring the same to a source without using any external work. This system, therefore,
violates the Clausius statement. (Note that in case of the combined system, out of the heat
Q1 + Q2 transferred to the source, Q1 is transferred from the source to the engine as a heat input to
the heat engine to produce work W (W = Q1) which is used to drive the heat pump. Therefore,
Q1 remains in a loop from the source - to the heat engine-to the heat pump and back-to the source
and the net effect is thus transfer of heat Q2 from the sink to the source).

Source at T2 Source at T1

Q1
Q1 + Q2 Q2

W = Q1
Heat Heat Heat
Engine Pump Pump

Q2 Q2

Sink at T2 Sink at T2

(a) Violation of the Kelvin-Planck statement (b) Final heat balance


(Engine violating the K-P statement)
Fig. 4.9
(2) Violation of Clausius statement results into violation of Kelvin-Planck statement :
Consider a heat pump which, operating in a cycle, absorbs heat Q2 from a sink and transfers the
same to a source with no other effect, i.e., without expenditure of work [Fig. 4.9 (a)]. This is
impossible as per the Clausius statement of the second law of thermodynamics. Therefore,
it violates the Clausius statement.
Let us assume that a heat engine that operates between the same two thermal reservoirs.
It absorbs heat Q1 from the source and rejects heat Q2 to the sink and produces work
W (W = Q1 – Q2). By combining the heat engine and the heat pump, the sink can be eliminated
as there is no net heat interaction with it because heat Q2 is rejected by the engine (to the sink)
and the same heat Q2 is taken up by the heat source (from the sink). So we see that the combined
system of heat engine and heat pump, i.e., the heat engine and the heat pump together, acts as a
heat engine [Fig. 4.10 (a)] which, operating in a cycle, absorbs heat Q1 – Q2 from a source and
converts it completely into work (W = Q1 – Q2) without rejecting heat to a sink (this engine
interacts with a single thermal reservoir and produces work) and therefore, this system violates
the Kelvin-Planck statement.
Chemical Engineering Thermodynamics - I 4.12 Second Law of Thermodynamics

Source at T2 Source at T1

Q2
Q1 Q1 - Q2

Heat Heat Heat


Engine W = Q1 - Q2 Engine W = Q1 - Q2
Pump

Q2 Q2

Sink at T2

(a) Violation of the Clausius statement (b) Final heat balance


(Heat pump violating Clausius statement)
Fig. 4.10
Carnot Cycle :
The complete conversion of heat into work is not possible. Therefore, the efficiency of a heat
engine cannot be 100%, i.e., no engine is 100% efficient. As an efficiency of 100% is not
possible we must know the maximum possible obtainable efficiency to compare the performance
of actual heat engines. We obtain a maximum efficiency only if the heat engine is operated in a
completely reversible manner, i.e., free of dissipative effect of any kind (since dissipative effects
are responsible for lowering the efficiency). A heat engine which operates on the Carnot cycle
(reversible cycle), a Carnot heat engine, gives us an idea regarding the maximum obtainable
efficiency. With this efficiency we can compare the efficiency of any actual heat engine to check
its performance for further improvements. Therefore, through the efficiency of a Carnot heat
engine, the ideal to be aimed at (in terms of efficiency) is known.
• A reversible cycle is a hypothetical cycle in which all processes making up the cycle are
reversible.
• Carnot cycle is a reversible cycle.
• A cyclic heat engine which operates on the Carnot cycle is called a Carnot heat engine.
The Carnot engine is impractical to build and operate since it comprises of reversible
processes. It serves as a standard of perfection with which we have to compare the
performance of any actual heat engine.
Sadi Carnot, a French engineer, in 1824 devised a classical reversible cycle in the study of
heat engines and the cycle was named after him in his honour Carnot cycle consists of two
reversible isothermal and two reversible adiabatic processes. It operates between a source at high
temperature T1 and a sink at low temperature T2.
Chemical Engineering Thermodynamics - I 4.13 Second Law of Thermodynamics

Consider a Carnot engine which operates between two thermal reservoirs (source and sink) at
temperatures T1 and T2. The Carnot heat engine consists of a cylinder fitted with a weightless
and frictionless piston (thus permitting reversible processes to be performed), consisting the
working substance which for simplicity may be taken as 1 mole of an ideal gas of constant heat
capacity. The engine receives heat from the source at T1 during part of its operating cycle and
rejects heat to the sink during another part of the cycle. No heat is exchanged between engine and
surroundings during the remaining parts of the cycle.
The Carnot cycle is shown in Fig. 4.11, where the pressure of the gas is plotted against its
volume in the cylinder. The cycle consists of four steps.
1. The cylinder containing the gas is kept in contact with the source at T1. The external
pressure is adjusted so that it is infinitesimally smaller than the gas pressure and the temperature
of the gas is infinitesimally smaller than that of the source. In this way, the gas is expanded
isothermally and reversibly until its volume increases to V2, by absorbing a quantity of heat Q1
from the source at T1. The path of the process is represented by the isothermal curve 1-2.
The heat added (or absorbed) to (or by) the system from the source in this process 1-2 is
V2
Q1 = RT1 ln V
1

Process 1-2 : Reversible isothermal heat addition at T1

Process 2-3 : Reversible adiabatic expansion from T1 to T2


1 Q1
T1 Process 3-4 : Reversible isothermal heat rejection at T2

2 Process 4-1 : Reversible adiabatic compression from T2 to T1


Pressure

4
T2
3
Q2

Volume

Fig. 4.11 : Pressure-volume changes for


the Carnot cycle
[Carnot cycle on a P-V diagram]
2. When point 2 is reached, the cylinder of the gas is thermally isolated from the source
and is surrounded by the non-conducting jacket so that the gas is expanded reversibly
(i.e., infinitesimally slowly) and adiabatically till its temperature falls from T1 to T2. Since the
work is done by the system in this process at the expense of its internal energy as the process
being adiabatic (no heat enters or leaves the system), the temperature of the system falls from T1
Chemical Engineering Thermodynamics - I 4.14 Second Law of Thermodynamics

to T2. The volume of the system increases from V2 to V3 and the pressure decreases from P2 to P3.
The path of the process is represented by the adiabatic expansion curve 2-3.
Since the process 2-3 is adiabatic, Q = 0 … no heat interaction with the surroundings.
3. The cylinder is now placed in thermal contact with the sink at T2, after removing the
non-conducting jacket. The gas is compressed reversibly and isothermally at T2, rejecting a
quantity of heat Q1 to the sink (the external pressure is maintained infinitesimally greater than the
gas pressure and the temperature of the gas infinitesimally greater than the temperature of the
sink). It is evident that the work is done on the system (as the process being compression).
Hence, heat produced is rejected/given out to the sink.
The heat rejected by the system to the sink in this process 3-4 is
V3
Q2 = RT2 ln V
4
4. At point 4, the cylinder is isolated from the sink at T2 and is again surrounded by the
non-conducting jacket. The gas is compressed reversibly and adiabatically until the initial state
(state-1 – V1, P1, T1) is restored/regained. Since this step being adiabatic compression, work is
done on the system and the temperature of the system increases from T2 to T1 (as the gas is
compressed adiabatically (Q = 0), internal energy of the gas increases due to work addition and
the temperature rises from T2 to T1).
Since the process 4-1 is adiabatic, Q = 0.
The net heat added to the system in the cycle is
V2 V3
Qnet = Q1 – Q2 = RT1 ln V – RT2 ln V
1 4
We have, W = Wnet = Qnet
V2 V3
W = Q1 – Q2 = RT1 ln V – RT1 ln V … (4.11)
1 4
Processes 2-3 and 4-1 are reversible adiabatic processes.
For these processes, V-T relationship is given by
TVγ–1 = constant
For process 2-3, it becomes
T1 (V2)γ–1 = T2 (V3)γ–1 … (4.12)
[since at state 2 : T = T1 and at state 3 : T = T2]
Similarly, for process 4-1 :
T2 (V4)γ–1 = T1 (V1)γ–1 … (4.13)
[since at state 4 : T = T2 and at state-1 : T = T1]
Rearranging Equations (4.12) and (4.13), we get
T1 V3γ–1
T2 = V2 … (4.14)

T1 V4γ–1
T2 = V1 … (4.15)
Chemical Engineering Thermodynamics - I 4.15 Second Law of Thermodynamics

Comparing Equation (4.14) with Equation (4.15), we get


V3 V4
V2 = V1
V2 V3
∴ V1 = V4 … (4.16)

Substituting Equation (4.16) into Equation (4.11), we get


W = Net work delivered = net work output = net work done by
the system
W = Q1 – Q2
V2 V2
= RT1 ln V – RT2 ln V
1 1

V2
W = R (T1 – T2) ln V … (4.17)
1

The net work delivered by the system during the cycle is represented by the enclosed area
(1-2-3-4-1) of the cycle.
The efficiency of a heat engine is defined as the ratio of the net work done in the cycle to the
heat absorbed/taken in from the source. The efficiency of the Carnot engine is
V2 V2
RT1 ln V – RT2 ln V
W Q1 – Q2 1 1
η = Q = Q = V2
1 1
RT1 ln V
1

W T1 – T2 T2
η = Q = T =1– T … (4.18)
1 1 1

Since T1 > T2, η < 1, where T1 and T2 are absolute temperatures (T1 and T2 are in K).
It is clear from the above equation that the efficiency depends upon the temperature of the
source and sink and not on the nature of the working substance. The efficiency can be increased
either by increasing the temperature of the source (T1) or decreasing the temperature of the sink
(T2) or by both. In practice, the sink temperature cannot be decreased below atmospheric
temperature since the atmosphere is used always as a sink, and so it is desirable that the source
temperature should be high. Because of this a high pressure steam is used for power production.
The Carnot cycle cannot be used in a practical engine because of the reasons given below :
(i) In the isothermal process of the cycle, the piston has to move very slowly in order to
provide a sufficient time for heat transfer and thus thereby temperature remains constant. In the
adiabatic process, the piston has to move as fast as possible. So that time available for heat
transfer is very short and thereby heat transfer from the system to the surroundings is negligible.
In this cycle, both isothermal and adiabatic processes take place during the same stroke of the
piston (isothermal expansion and adiabatic expansion). Therefore, the piston has to move very
Chemical Engineering Thermodynamics - I 4.16 Second Law of Thermodynamics

slowly for part of the stroke and it has to move very fast during the remaining part of the stroke.
This variation of speed of the piston during the same stroke is not possible kinematically. The
adiabatic process demands perfect insulation of the cylinder walls.
(ii) It is not possible to perform a frictionless process.
(iii) It is not possible to transfer heat without a finite temperature difference.
A heat engine which operates on the Carnot cycle is called a Carnot engine.
The Carnot cycle for a steady flow system is shown in Fig. 4.12. Heat Q1 is added to the
system reversibly and isothermally at T2 in a heat exchanger (A), work WT is done by the system
reversibly and adiabatically in a turbine (B), heat Q2 is removed from the system reversibly and
isothermally at T2 in a heat exchanger (C) and finally, work WP is done on the system reversibly
and adiabatically by a pump (D), thus completing the cycle. In order to meet the conditions of the
Carnot cycle, friction in the pipelines (through which the working fluid flows) and heat transfer
through the pipelines must be absent.

Source at T1

Q1

Heat exchanger
(A) at T1

WP Pump Turbine WT
(D) (B)

System
Heat exchanger boundary
(C) at T2

Q2

Sink at T2

Fig. 4.12 : Carnot heat engine - steady flow system


The net work transfer / net work delivered by the system = W = Wnet
= WT – Wp = Q1 – Q2
Carnot theorems / Carnot principles :
Carnot theorem 1 :
• No heat engine operating between two given constant temperature thermal reservoirs
can be more efficient than a reversible heat engine operating between the same thermal
reservoirs.
• No heat engine operating between two given constant temperature thermal reservoirs
has a higher efficiency than a reversible heat engine.
Chemical Engineering Thermodynamics - I 4.17 Second Law of Thermodynamics

Carnot theorem 2 :
• All reversible heat engines operating between the same two given constant temperature
thermal reservoirs have the same efficiency.
• All reversible heat engines operating between two given temperatures (operating
between the same two temperatures) have the same efficiency.
In other words, the efficiency of a reversible heat engine is independent of the nature of the
working substance, but depends on the temperatures of the thermal reservoirs between which it
operates.
Maximum efficiency of heat engine :
The efficiency of a reversible heat engine operating between two given temperatures is the
maximum possible that can be expected for any heat engine, for those temperatures.
This can be understand by considering the Carnot engine mentioned earlier. The work terms
in the adiabatic processes 2-3 and 4-1 cancel one another. Therefore, the work done in the Carnot
cycle is that involved in the isothermal expansion and isothermal compression processes
(processes 1-2 and 4-1 respectively).
We know that in an isothermal expansion, the work done by the system (work obtained) is a
maximum if the process of expansion is reversible. Also we know that in an isothermal
compression, the work done on the system is a minimum if the process of compression is a
minimum if the process of compression is reversible (maximum work is obtained in a reversible
isothermal expansion and minimum work is required in a reversible isothermal compression).
Consequently, in the Carnot cycle the work obtained, i.e., the work done by the system is the
maximum possible for the isothermal expansion at T1 from 1 to 2, whereas the work done on the
system at T2 is the minimum work required for the compression from 3 to 4 since both the
processes are reversible. It is clear, therefore, that the total work done by the system, i.e., the net
work done by the system in the cycle is the maximum for the given conditions. Since the
efficiency is directly proportional to the net work done by the system, the efficiency of a
reversible heat engine is the maximum possible for the given temperatures.
Thermodynamic temperature scale :
Use of the efficiency of a reversible engine as the basis of a temperature scale was suggested
by Lord Kelvin in 1848.
In case of a reversible heat engine operating between two given thermal reservoirs, the
temperature of each reservoir on the thermodynamic (Kelvin) scale is defined as proportional to
the quantity of heat transferred to or from it in a reversible cycle. If Q1 is the heat transferred
from the thermal reservoir at the higher temperature and Q1 is the heat transferred to the thermal
reservoir at the lower temperature, then the respective temperatures on the thermodynamic
temperature scales θ1 and θ2, without taking into account the signs for heat quantities, are given
by
θ1 Q1
= Q … (4.19)
θ2 2
Chemical Engineering Thermodynamics - I 4.18 Second Law of Thermodynamics

If Q1 and Q2 are measured, then their ratio is equal to the ratio of the thermodynamic
temperatures of the two thermal reservoirs (source and sink) and this ratio of the two
temperatures is independent of a thermometric substance. If we operate a heat engine between a
thermal reservoir at θ1 and another thermal reservoir at θ2 = 273.16 K (i.e., at the triple point of
water) and measure Q1 and Q2, then we can determine the temperature θ1.
Inverting each side of the above equation and then subtracting each inverted side from unity,
we get
θ2 Q2
1– = 1– Q
θ1 1

θ1 – θ2 Q1 – Q2
∴ = Q1
θ1
Q1 – Q2 θ1 – θ2
Q1 =
θ1
We know that the net work done by a heat engine is given by
W = Q1 – Q2 = heat absorbed – heat rejected
W Q1 – Q2 θ1 – θ2
∴ Q1 = Q1 = θ1
W
We know that Q1 = η. Therefore,
W Q2 θ2
η = Q =1– Q = 1– … (4.20)
1 1 θ 1
This equation defines the efficiency of the reversible engine in terms of the thermodynamic
temperatures.
We have already seen that the efficiency of a Carnot engine (it is a reversible engine) which
uses an ideal gas as the working medium is given by
T1 – T2 T2
η = T =1– T … (4.20)
1 1

The temperatures T1 and T2 are measured by ideal gas thermometers or ideal gas temperature
scale.
Comparing Equations (4.19) and (4.20), we get
Q2 T2 θ2
Q1 = T1 = θ1 … (4.21)

That is, the ideal gas temperature scale and the thermodynamic temperature scale are the
same or identical. As a result of the identity of the ideal gas scale and the thermodynamic
temperature scale, temperature on the former scale, like those on the latter scale, may be regarded
as absolute. These are independent of the thermometric substance and measured in degrees
Kelvin (K). Having proved the similarity between these two temperature scales, we use T to
denote temperature. As a result of this, we can write
Q2 T1
Q1 = T2
Chemical Engineering Thermodynamics - I 4.19 Second Law of Thermodynamics

COP of a Carnot heat pump/refrigerator :


We have already seen that the efficiency of a Carnot engine is given by
T1 – T2
η = T
1

As this is a reversible engine, any other reversible engine operating between the same two
temperatures should also have the same efficiency. Therefore, the efficiency of a reversible
engine operating between temperatures T1 and T2 is given by
Q1 – Q2 T1 – T2
η = ηmax = Q1 = T1
As all the processes of the Carnot cycle are reversible, each individual process can be
reversed and carried out in reverse order. If the Carnot cycle is reversed (4-3-2-1-4), it becomes
a Carnot heat pump or refrigerator (reversed Carnot heat engine). Here, the quantities Q1, Q2 and
W remain the same in magnitude but their directions are reversed. The Carnot heat pump or
refrigerator takes heat Q2 from a low temperature body (at T2), discharges heat Q1 to a high
temperature body (at T1) and receives net work W.
The purpose of a refrigerator is to extract heat from a body at low temperature (i.e., cooling).
The coefficient of performance of a Carnot refrigerator is given by
Q2 Q2 Heat extracted from the low T body
[COP]R = W = Q – Q = Work required/input
1 2
1
= Q
1
Q2 – 1
Q1 T1
We know that : Q2 = T2
1
Therefore, [COP]R = T
1
T2 – 1
T2
[COP]R = T – T … (4.22)
1 2

The purpose of a heat pump is to supply or deliver heat to a high temperature body
(i.e., heating instead of cooling). The coefficient of performance of a Carnot heat pump is given
by
Q1 Heat delivered to the high T body
[COP]HP = W = Work required
Q1
= Q –Q
1 2
Dividing both numerator and denominator by Q2 gives
Q1/Q2
[COP]HP = Q
1
Q2 – 1
Chemical Engineering Thermodynamics - I 4.20 Second Law of Thermodynamics

Q1 T1
We know that : Q2 = T2
T1/T2
∴ [COP]HP = T
1
T2 – 1
T1
[COP]HP = T – T … (4.23)
1 2

Two reversible adiabatic paths cannot intersect each other :


This statement can be proved by contradiction. That is, by assuming that such paths intersect
each other.
Assume that two reversible adiabatic paths AC and BC intersect each other at point C as
shown in Fig. 4.13, where they are plotted on P-V coordinates.

A
Reversible isothermal
process

B
Pressure

Reversible adiabatic
processes

Volume
Fig. 4.13 : Assumption of two reversible adiabatics intersecting each other
Let a reversible isothermal path AB intersects the reversible adiabatic paths at A and B.
These three reversible processes AB, BC and CA together constitute a cycle. The area enclosed
by ABC (area of the cycle) represents the net work output in the cycle. A heat engine operating
on this cycle receives heat during the isothermal process AB by interacting with a single thermal
reservoir and delivers work equal to area of the cycle, without rejecting heat to any other thermal
reservoir. This violates the Kelvin-Planck statement of the second law of thermodynamics
(this forms a PMM2 which is impossible). Therefore, our assumption of the intersection of the
two reversible adiabatic paths (adiabatics) is wrong. Hence, it is proved that two reversible
adiabatic paths cannot intersect each other. It means that only one reversible adiabatic can pass
through a given point.
(In Fig. 4.13 two reversible adiabatics intersect at point C, i.e., both of them have a common
state C. Since this cannot be possible, it is also stated that two reversible adiabatic processes
cannot have a common state).
Chemical Engineering Thermodynamics - I 4.21 Second Law of Thermodynamics

Reversible adiabatic lines are steeper than isothermal lines on a P-V diagram :
The reversible adiabatic and isothermal processes are represented on a P-V diagram as
shown in Fig. 4.14.

Pressure Adiabatic
process
Isothermal
process

Volume
Fig. 4.14 : Reversible adiabatics steeper than isotherms
For an ideal gas as the working substance, these processes are represented by the relations
PV = C … isothermal process
PVγ = C … adiabatic process
The slope of the isothermal line is given by
dP P
dV = – V … (4.24)
We have : PV = C
Differentiating with respect to V gives
dP
P × 1 + V dV = 0
dP dP P
V dV = – P ∴ dV = – V
The slope of the adiabatic line is given by
dP P
dV = – γ V
We have : PVγ = C
Differentiating with respect to V gives
dP
P · γ Vγ–1 + Vγ dV = 0

dP
Vγ dV = – γ P Vγ–1

dP Vγ–1
= – γ P = – γ · P Vγ–1–γ = – γ PV–1
dV Vγ


dP P
dV = – γ V … (4.25)
Chemical Engineering Thermodynamics - I 4.22 Second Law of Thermodynamics

Comparing Equations (4.24) and (4.25), we can write

 dP   dP 
dVadiabatic > dVisothermal, since γ > 1

(A line having a higher value of the slope is steeper than a line having a lower value of the
slope.)

Therefore, a reversible adiabatic line is always steeper than a reversible isothermal line. This
also indicates that these two lines cannot touch each other without intersection.

Any reversible process can be replaced by two reversible processes and one isothermal
process :

Consider that a system undergoes a reversible process from state A to B (Fig. 4.15). Let this
reversible process AB be replaced by a reversible adiabatic process AC and a reversible
isothermal process CD followed by another reversible adiabatic process CD (so we ultimately
reach the state B : A → C → D → B). The isothermal process CD is performed in such a way
that when it is plotted on P-V coordinates as shown in Fig. 4.15, the area enclosed by ACO is
equal to the area enclosed by ODB, so that the sum of the areas under the curves AC, CD and DB
is equal to the area under the curve AB (i.e., the area under the curve A-C-D-B is equal to the
area under the curve AB).

Adiabatic lines

Isothermal line
C

Pressure O Area ACO = Area ODB


A
D

Volume

Fig. 4.15 : Reversible process replaced by two reversible adiabatic processes and
one reversible isothermal process
The area under the curve on a P-V diagram represents the work done during the process.
∴ WA-B = WA-C-D-B
Chemical Engineering Thermodynamics - I 4.23 Second Law of Thermodynamics

Applying the first law of thermodynamics for the process AB gives


QA-B = (UB – UA) + WA-B, [∆U = Q – W ⇒ Q = ∆U + W] … (4.26)
Applying the first law of thermodynamics for the process ACDB gives
QA-C-D-B = (UB – UA) + WA-C-D-B … (4.27)
Since WA-C-D-B = WA-B, from Equations (4.26) and (4.27), we can write
QA-B = QA-C-D-B
i.e., QA-B = QAC + QCD + QDB … (4.28)
Hence, heat transferred during the reversible process AB is equal to the heat transferred
during the combination of processes AC, CD and DB.
In an adiabatic process, Q = 0. Therefore, for the two adiabatic processes AC and DB,
QA-C = 0 and QD-B = 0.
Since QA-C = 0 and QD-B = 0, Equation (4.28) becomes
QA-B = QC-D
Heat transferred in the process AB is equal to the heat transferred in the isothermal
process CB.
Thus, any reversible process may be replaced by a reversible adiabatic process and a
reversible isothermal process followed by a reversible adiabatic process (i.e., by two reversible
adiabatic processes and one reversible process) between the same end states such that the heat
transferred during the reversible isothermal process is the same as transferred during the original
reversible process. Since a process is a path followed by a system in reaching the given final state
from the initial state, it may be stated that any reversible path may be replaced by two reversible
adiabatic paths and one reversible isothermal path.
Based on this we can approximate a reversible cycle into a large number of Carnot cycles.
Clausius theorem and Clausius Inequality :
We know that any reversible path can be replaced by two reversible adiabatic paths and one
reversible isothermal path between the same end states. The sequence being a reversible
adiabatic path followed by a reversible isothermal path which in turn is followed by a reversible
adiabatic path.
Consider a reversible cycle, which is represented by a closed curve as shown in Fig. 4.16.
Divide this cycle into a large number of strips by drawing reversible adiabatic lines/curves. Since
two adiabatic lines cannot intersect, they may be drawn as close as possible. A few of these
curves are shown on the figure as long continuous lines. Connect adjacent adiabatic curves by
short reversible isothermal lines at the top and bottom (such that the areas of the triangles formed
by any two adjacent adiabatic lines, corresponding isothermal line and reversible path are the
same).
Chemical Engineering Thermodynamics - I 4.24 Second Law of Thermodynamics

Each pair of adjacent adiabatic lines and their isothermal connecting lines represent a Carnot
cycle. In this way, the original cycle is subdivided into a large number of Carnot cycles. Note that
each adiabatic line forms a part of two adjacent Carnot cycles. If the adiabatic lines are placed
very close to each other, the number of Carnot cycles is very large and the isothermal lines
approximate the original cycle. Therefore, the original reversible cycles may be regarded as
equivalent to the contribution of an infinite number of small Carnot cycles.

Q1
T1

1
A B
B A 2
Area A12 = Area 23B
3

Pressure

4
C 5 D
D Area D56 = Area 54C
C 6
T2
Q2
Volume
Fig. 4.16 : A reversible cycle subdivided into a large number of Carnot cycles
Each Carnot cycle has its own pair of isotherms T1 and T2 and associated heat quantities
Q1 and Q2, which are indicated on the figure for a representative cycle.
We know that for a Carnot cycle (reversible cycle),
Q2 T2
Q1 = T1
Q2 Q1
or T2 = T1
where Q1 is the heat absorbed at T1 and Q2 is the heat rejected by the system at T2. Q1 is positive
and Q2 is negative as per the sign conventions adopted. Therefore, using these sign conventions,
the above equation becomes
– Q2 Q1
T2 = T1
Q1 Q2
i.e., T1 + T2 = 0 … (4.28 a)
Q
or ∑ T = 0
Chemical Engineering Thermodynamics - I 4.25 Second Law of Thermodynamics

An infinite number of infinitesimal Carnot cycles are required to duplicate the original cycle.
Therefore, when the adiabatic lines very closely spaced, the original cycle consists of an infinite
number of infinitesimal Carnot cycles. For an infinitesimal cycle, isothermal steps are
infinitesimal and thus the heat quantities become dQ1 and dQ2.
For an infinitesimal Carnot cycle, Equation (4.28 a) becomes
dQ1 dQ2
T1 + T2 = 0 … (4.29)

Similar equations can be written for all the infinitesimal Carnot cycles.
Therefore, for the original cycle consisting of an infinite number of infinitesimal Carnot
cycle, we can write
dQ1 dQ2 dQ3 dQ4
T1 + T2 + T3 + T4 + … = 0
dQrev
or ⌠

O T = 0 … (for a reversible cycle) … (4.30)

where the circle in the integral sign signifies that integration is over a complete cycle.
The subscript 'rev' signifies that the equation is valid only for reversible cycles.
Equation (4.30) is known as Clausius theorem. It states that whenever a system undergoes a
reversible cycle, the cyclic integral of dQ/T is zero (or simply the cyclic integral of dQ/T for a
reversible cycle is zero).
The efficiency of a reversible cycle is given by
Q2 T2
η = 1–Q =1– T
1 1

The efficiency of an irreversible cycle is less than that of a reversible cycle.


1 – Q2  T 2
 Q < 1 – T 
 
1 irreversible  1reversible

The above statement is true if


Q2 T2
Q1 > T1
Q2 Q1
T2 > T1
Adopting the sign convention for heat quantities (added : +ve, rejected : –ve), the above
equation becomes
– Q2 Q1
T2 > T1
Q1 Q2
i.e., T1 + T2 < 0
Q
or ∑T < 0
An irreversible cycle can be divided into a large number of infinitesimal cycles involving
infinitesimally small heat interactions (e.g., dQ1 and dQ2).
Chemical Engineering Thermodynamics - I 4.26 Second Law of Thermodynamics

For an infinitesimal cycle, we can write,


dQ1 dQ2
T1 + T2 < 0
Therefore, for an irreversible cycle comprising of a large number of infinitesimal irreversible
cycles, we can write
dQ

⌡O T < 0 [for an irreversible cycle] … (4.31)

Combining Equation (4.30), which is applicable for any reversible cycle with Equation
(4.31), which is applicable for any irreversible cycle, we get
dQ

⌡O T ≤ 0 … (4.32)

This equation is known as the inequality of Clausius or Clausius inequality. It states that the
cyclic integral of dQ/T is less than or equal to zero depending on whether a given cycle is
reversible or irreversible.
The equality sign holds good for reversible cycles and the inequality sign holds good for
irreversible cycles.


O dQ/T is zero for a reversible cycle which is highly impractical to achieve in practice.
Therefore, for all practical cycle,
dQ


O T < 0

Equation (4.32) provides the criterion for the reversibility of a cycle or cyclic process.
dQ
If ⌠

O T = 0, the cycle is reversible

dQ


O T < 0, the cycle is irreversible and possible

dQ


O T > 0, the cycle is impossible

Entropy :
Potential energy is a high grade energy, whereas heat energy is a low grade energy. This is
due to the fact that a higher portion of the potential energy can be converted into work (the most
desirable form of energy) than that of the heat energy.
1 kJ of heat energy available at 1000 K may be converted into 0.70 kJ of work, but the same
quantity of heat energy available at 500 K may not be able to produce even 0.5 kJ of work. Thus
is, more work is obtained from the heat energy available at high temperature than that can be
obtained from the same amount of heat energy available at low temperature. This indicates that it
is not the quantity but the quality of energy that is more important.
Chemical Engineering Thermodynamics - I 4.27 Second Law of Thermodynamics

All processes occurring in nature are irreversible. Some processes are highly irreversible,
e.g., stopping of a car by applying brakes, whereas others are slightly irreversible, e.g.,
expansion/compression of gases in a cylinder.
Therefore, it is necessary to establish some form of measure to decide the quality of energy
and irreversibility of a process in a mathematical form. To measure the quality of energy or to
measure the irreversibility of a process, Clausius introduced by the concept of an important
property called entropy. Entropy is an outcome of the second law of thermodynamics and used to
predict the feasibility of a process.
Suppose a system is taken from the initial state 1 to the final state 2 by following a reversible
path A (Fig. 4.17). The system is restored to its initial state 1 by following another reversible
path B. Then the two paths A and B together constitute a reversible cycle. The reversible cycle is
1-A-2-B-1.
The Clausius inequality for this cycle can be written as
dQ


O T = 0
(AB)

This cyclic integral can be replaced by the sum of two integrals, one each for paths A and
B, as
2 1
dQ dQ

⌡ T + ⌠
⌡ T = 0
1 (A) 2 (B)

2 1
dQ dQ

⌡ T = – ⌠
⌡ T
1 (A) 2 (B)

2 2
dQ dQ
∴ ⌠
⌡ T = ⌠
⌡ T … (4.33)
1 (A) 1 (B)

Fig. 4.17 : Two reversible cycles on P-V diagram


Chemical Engineering Thermodynamics - I 4.28 Second Law of Thermodynamics

If the system restored to state 1 from state 2 by following a reversible path C, then the two
paths A and C together constitute another reversible cycle. The reversible cycle is 1-A-2-C-1.
The Clausius inequality for this cycle can be written as
dQ


O T = 0
(AC)

This cyclic integral can be replaced by the sum of two integrals, one for the path A and the
other for the path C, as
2 1
dQ dQ

⌡ T + ⌠
⌡ T = 0
1 (A) 2 (C)
2 1
dQ dQ

⌡ T = – ⌠
⌡ T
1 (A) 2 (C)
2 2
dQ dQ
∴ ⌠
⌡ T = ⌠
⌡ T … (4.34)
1 (A) 1 (C)

From Equations (4.33) and (4.34), we get


2 2 2
dQ dQ dQ
⌠ ⌠ T = ⌡
⌡ T = ⌡ ⌠ T … (4.35)
1 (A) 1 (B) 1 (C)

That is, the value of ⌠


⌡ dQ/T of the system in going from state 1 to state 2 is the same
irrespective of the path followed by the system (A, B or C) provided that they are reversible - the
value of ⌠
⌡ dQ/T is the same over all reversible paths and so it does not depend on the path
followed by the system.

Hence, the expression ⌠


⌡ dQ/T, which is independent of the path followed, is an exact
differential, represents a property of the system. This property is called entropy. It is denoted by
S and defined mathematically as
dQrev
dS = T
2
dQrev
or ∆S = S2 – S1 = ⌠
⌡ T … (4.36)
1

The subscript rev indicates that dQ/T is to be evaluated along a reversible path only.
The entropy is a state function. The change in the entropy of a system, being a property of
the system, depends only on the initial and final states of the system and not on the path followed
by the system. Therefore, the change in entropy between two given states is the same whether the
process is reversible or irreversible.
Chemical Engineering Thermodynamics - I 4.29 Second Law of Thermodynamics

Characteristics of entropy :
• The change in entropy of a system undergoing a reversible process between states 1 and
2 is given by the relation
2
dQrev
∆S = ⌠
⌡ T … (4.37)
1

The entropy change of the system may be positive, negative or zero.


• In case of a reversible process, the change in entropy of the surroundings is equal in
magnitude, but opposite in sign, to that of the system. That is, the net entropy change
(or the sum of the entropy changes) of a system and its surroundings is zero for any
reversible process.
• Since the entropy change is the same over all reversible paths, it may be evaluated over
any convenient reversible path.
• Entropy is a state function/point function and its differential is exact. It is a property of
the system.
• Since entropy is a state function, the entropy change of system for a finite change of
state is invariably the same whether the system follows a reversible process or an
irreversible process. Therefore, for an irreversible process, the entropy change of a
system is evaluated over a reversible path (using the relation for a reversible process)
even though the process is irreversible. The entropy change of a system undergoing an
irreversible process may be positive, negative or zero.
• The sum of the entropy changes of the system and its surroundings is always greater
than zero for an irreversible process.
• In all processes in nature (i.e., in real processes), entropy is not conserved but created.
The change in entropy of a system between any given two states is the integral of the
reversible heats divided by the absolute temperature, over a reversible path. The change in
entropy of a system undergoing a reversible process may be positive, negative or zero. For
example, in the reversible isothermal expansion of a gas, heat is added to the gas and therefore,
the entropy change of the gas is positive, whereas in the reversible isothermal compression, heat
is removed from the gas and therefore, the entropy change of the gas is negative. In an adiabatic
reversible expansion/compression of gas, no heat is added or removed and hence the entropy
change of the gas is zero.
Using Equation (4.37) we can estimate the changes in entropy and absolute values of the
entropy are determined using the third law of thermodynamics.
Units of entropy :
In the SI system, the units of entropy are J/K and these are represented by EU, indicating
entropy units. The units of specific entropy are J/(kg.K).
If the system contains m kg of a substance, then
Total entropy change in J/K = m, kg × specific entropy, J/(kg.K)
Chemical Engineering Thermodynamics - I 4.30 Second Law of Thermodynamics

Physical significance of entropy :


• The entropy of a system is regarded as a measure of that portion of the total energy of
the system which is unavailable to perform useful work.

• The entropy of a system is a measure of the degree of disorder or randomness of the


system.

• When a system does some work then we say that the potential ability of the system
decreases that is the potential inability of the system increases or entropy increases.

• In all processes in nature (in real processes), entropy is not conserved but created. The
entropy generated in a process is regarded as a measure of the irreversibility of the
process.

Principle of entropy increase :


Consider two different cycles undergone by a system as shown in Fig. 4.18. The paths A and
B are reversible, while the path C is irreversible. The cycle 1A2B1 is a reversible cycle, in which
the system is taken from the initial state 1 to the final state 2 by following the reversible path A
and is then restored to the initial state 1 by following the reversible path B. The cycle 1A2C1 is
an irreversible cycle in which the system is restored to the initial state by following the
irreversible path C.

A
C
v.

Pressure
Re

v.
I rre B
v.
1 Re

Volume

Fig. 4.18 : A reversible and an irreversible cycle undergone by a system

Applying the Clausius inequality to the reversible cycle 1A2B1, we get


2 1
dQ dQ dQ

O T = ⌠
⌡ ⌡ T + ⌠
⌡ T =0
(AB) 1 (A) 2 (B)

2 1
dQ dQ
Or ⌠
⌡ T = – ⌠
⌡ T … (4.38)
1 (A) 2 (B)
Chemical Engineering Thermodynamics - I 4.31 Second Law of Thermodynamics

Applying the Clausius inequality to the irreversible cycle 1A2C1, we get


2 1
dQ dQ dQ

O T = ⌠
⌡ ⌡ T + ⌠
⌡ T < 0 … (4.39)
(AC) 1 (A) 2 (C)

Substituting Equation (4.38) in Equation (439), we get


1 1
dQ dQ
– ⌠
⌡ T + ⌠
⌡ T < 0
2 (B) 2 (C)
1 1
dQ dQ
i.e., ⌠
⌡ T – ⌠
⌡ T > 0
2 (B) 2 (C)
1 1
dQ dQ
Or ⌠
⌡ T > ⌠
⌡ T … (4.40)
2 (B) 2 (C)

Since the path B is reversible,


1 1
dQ

⌡ dS = ⌠
⌡ T … (4.41)
2 (B) 2 (B)

As entropy being a property, entropy changes from 2 to 1 along the paths B and C would be
the same. Therefore,
1 1
⌠ dS = ⌡
⌡ ⌠ dS … (4.42)
2 (B) 2 (C)

Combining Equations (4.40), (4.41) and (4.42) gives


1 1 1
dQ
⌠ dS = ⌡
⌡ ⌠ dS > ⌡
⌠ T
2 (B) 2 (C) 2 (C)
Therefore, for any irreversible process, we can write
dQ
dS > T … (4.43)

That is, the entropy change of a system is greater than dQ/T, where dQ represents the heat
interaction along the irreversible path. The entropy change of a system for reversible process is
given by
dQrev
dS = T … (4.44)

Therefore, by combining Equations (4.43) and (4.44), we can write for any process,
dQ
dS ≥ T … for an infinitesimal change/process
2
dQ
or ∆S = S2 – S1 ≥ ⌠
⌡ T … for definite change … (4.45)
1

where the equality sign holds good for a reversible process and inequality sign holds good
(applicable) for an irreversible process.
Chemical Engineering Thermodynamics - I 4.32 Second Law of Thermodynamics

For an isolated system undergoing some change of state, dQ = 0 (i.e., no heat interaction
with the surroundings) and therefore, Equation (4.45) gives
dS ≥ 0 or ∆S ≥ 0 … for an isolated system … (4.46)
The equality sign holds good for a reversible process and the inequality sign holds good for
an irreversible process (a naturally occurring process or a real process).
(∆S = 0 or S = constant … reversible process
∆S > 0 … irreversible process)
That is, the entropy of an isolated system always increases (in any real process) or remains
constant (in a reversible process), but can never decrease. This is known as the principle of
entropy increase or simply entropy principle.
Equation (4.46) is the mathematical statement of the second law of thermodynamics.
The combination of a system and its surroundings can be referred to as an isolated system
since this system can be formed by including the system and its surroundings within a single
boundary. Since the universe is a combination of the system and its surroundings, it is an isolated
system. Therefore,
∆Suniv ≥ 0
or ∆Ssys + ∆Ssurr ≥ 0 … (4.47)
The principle of entropy increase gives us an idea regarding the feasibility of a process. The
feasibility criterion in terms of entropy change is
∆S > 0 : Process is feasible and irreversible
∆S = 0 : Process is feasible and reversible (hypothetical)
∆ S < 0 : Process is not feasible
Applications of the principle of increase of entropy :
The combination of a system and its surroundings is referred to as an isolated system. Thus,
the universe comprising of the system and surroundings is an isolated system.
Every irreversible process is accompanied by increase of entropy of the universe. The
increase of entropy of the universe is a measure of the extent of irreversibility of the process
undergone by the system. The higher the increase in entropy of the universe, the higher will be
the irreversibility of the process.
• Heat transfer through a finite temperature difference :
Let Q be the amount of heat transferred from a heat source at T1 to a heat sink at T2 (T1 > T2).
The change in entropy of the heat source is – Q/T1. It is negative because heat Q flows out of the
source. The entropy change for the sink is Q/T2. It is positive because heat flows into the sink.
Therefore, for the isolated system comprising of the heat source and sink
∆Stotal = ∆Suniv. = ∆Ssource + ∆Ssink
–Q Q
= T +T
1 2

∆Suniv = Q  T T 
T1 – T2
… (4.48)
 1 2
Chemical Engineering Thermodynamics - I 4.33 Second Law of Thermodynamics

Since T1 > T2 and Q is positive, ∆Suniv is positive and the process of heat transfer with a
finite temperature difference between the source and sink is irreversible and possible. If T1 = T2,
∆Suniv is zero and the process is reversible. If T1 < T2, ∆Suniv is negative and the process is
impossible.
Adiabatic mixing of two fluids :
Consider an adiabatic enclosure which is divided into two compartments with the help of a
partition. One compartment contains a fluid-1 of mass m1, heat capacity Cp1 and temperature T1
and the other compartment contains a fluid-2 of mass m2, heat capacity Cp2 and temperature T2.
When the partition is removed, the two fluids mix together. Let Tf be the final temperature
attained by both the fluids at equilibrium such that T2 < Tf < T1. Since the system is isolated,
being adiabatic mixing, heat interaction is confined to the two fluids.
Heat given out by fluid-1 = Heat gained by fluid-2
m1 Cp1 (T1 – Tf) = m2 Cp2 (Tf – T2)
m1 Cp1 T1 – m1 Cp1 Tf = m2 Cp2 Tf – m2 Cp2 T2
m1 Cp1 T1 + m2 Cp2 T2
∴ Tf = m1 Cp1 + m2 Cp2
The entropy change for the fluid-1 is given by
Tf f T
dQrev mf Cp1 dT Tf
∆S1 = ⌠
⌡ T = ⌠ ⌡ T = m1 Cp1 ln T
1
T1 T1

Since T1 > Tf , ∆S1 will be negative.


The entropy change for the fluid-2 is
Tf
m2 Cp2 dT Tf
∆S2 = ⌠
⌡ T = m2 Cp2 ln T
2
T2

Since T2 < Tf, ∆S2 will be positive.


The total entropy change of this isolated system due to mixing process is the sum of the
individual entropy changes since entropy is an additive property.
∆S (or ∆Sm) = ∆Suniv = ∆S1 + ∆S2
Tf Tf
∆Suniv = m1 Cp1 ln T + m2 Cp2 ln T … (4.49)
1 2

If m1 = m2 = m and Cp1 = Cp2 = Cp, then Equation (4.49) becomes


 Tf Tf   Tf Tf
∆Suniv = mCp ln T + ln T  = mCp ln T × T 
 1 2   1 2
2
Tf
∆S = ∆Suniv = m Cp ln T T
1 2
Chemical Engineering Thermodynamics - I 4.34 Second Law of Thermodynamics

m1 Cp1 T1 + m2 Cp2 T2
We have, Tf = m1 Cp1 + m2 Cp2
m Cp T1 + m Cp T2 T1 + T2
= m Cp + m Cp = 2
2
Tf  Tf 2
We have, ∆Suniv = m Cp ln T T = m Cp ln  
1 2  T 1T 2
Tf
= 2m Cp ln
T 1T 2

Substituting for Tf,


(T1 + T2)/2
∆S (or ∆Sm) = ∆Suniv = 2m Cp ln … (4.50)
T1 T2

Since the arithmetic mean of any two numbers is always greater than their geometric mean,
∆Suniv is always positive. Therefore, the mixing of two fluids is an irreversible process.

With the help of Equation (4.49) we can calculate the change in entropy occurring in the
process of mixing of two fluids. This equation can also be used to calculate the change in entropy
during an adiabatic mixing of two substances at different temperatures. For example, quenching
of metallic bodies in liquids.

Isothermal mixing of ideal gases :


Consider a number of non-interacting ideal gases placed in a vessel of volume V in the
separate compartments formed by partitions. Let ni be the number of moles of any gas and Vi be
the volume it occupies. The total entropy of the system before mixing is the sum of entropies of
the gases in the separate compartments.
S1 = ∑ ni [Cv ln T + R ln Vi + Ci] … (4.51)

Since the gases are at the same temperature before and after mixing - isothermal mixing.
dT dV
[dS = Cv T + R V [Equation (4.64 a)] ⇒ dS = Cv d ln T + R d ln V ⇒ S = Cv ln T + R ln V + C]

where C is the constant of integration.


If the partitions are removed, the gases will get mixed.
Each gas now occupies the volume V of the vessel (which is the total volume of the system).
The total energy of the system after mixing is given by
S2 = ∑ ni [Cv ln T + R ln V + Ci]
Chemical Engineering Thermodynamics - I 4.35 Second Law of Thermodynamics

If the pressure of each gas before mixing is the same as the total pressure of the mixing of
ideal gases, then
Individual volume of any gas Individual moles of any gas
Total volume of gas mixture = Total moles of gas mixture
Vi ni ni
V = ∑ ni = n = xi
where xi is the molefraction of the given gas in the gas mixture.
∴ V i = xi V … (4.52)
Substituting Equation (4.51) into Equation (4.50), we get
S1 = ∑ ni [Cv ln T + R ln xi + R ln V + Ci] … (4.53)
The increase of entropy of the system resulting from the mixing of gases, i.e., the entropy of
mixing of the gases is given by
∆S = ∆Sm = S2 – S1
∆S = ∑ ni [Cv ln T + R ln V + Ci] – ∑ ni [Cv ln T + R ln xi + R ln V + Ci]
∆S = – R ∑ ni ln xi … (4.54)
The entropy of mixing of 1 mole of the mixture of ideal gases is obtained by dividing
Equation (4.54) by the total number of moles, n.
ni
∆S = – R ∑ n ln xi

∆S = ∆Sm = – R ∑ xi ln xi … (4.55)
For mixing of two ideal gases,
∆S = ∆Sm = – R [x1 ln x1 + x2 ln x2] … (4.56)
The mole fraction xi of any gas in a gas mixture is less than unity, so its logarithm, i.e., ln xi
is negative. Therefore, ∆S given by Equation (4.55) is always positive. Hence, the mixing of
gases, e.g., by diffusion, is always accompanied by an increase of entropy and thus, it is an
irreversible process.
Temperature-Entropy plot :
The differential change in entropy dS due to reversible heat transfer dQ at temperature T is
given by the expression
dQrev
dS = T

In an adiabatic process, no heat enters or leaves the system [i.e., dQrev = 0]. Therefore, for a
reversible adiabatic process,
dS = 0 Therefore, S = constant
or ∆S = S2 – S1 = 0, i.e., S2 = S1 = S … for finite change
Chemical Engineering Thermodynamics - I 4.36 Second Law of Thermodynamics

A reversible adiabatic process on a T-S diagram is represented by a straight line parallel to


the T-axis, as shown in Fig. 4.19.
A reversible adiabatic process is called as an isentropic process since S is constant
[isentropic means same entropy in Greek].

T
1

Fig. 4.19 : Reversible adiabatic process on T-S diagram


dQrev
We have, dS = T
dQrev = T dS

For reversible isothermal heat transfer, T = constant. Therefore,


Qrev = Q = T [S2 – S1]

A reversible isothermal process on a T-S diagram is represented by a straight line parallel to


the S-axis, as shown in Fig. 4.20.

1 2

Q
T

Fig. 4.20 : T-S diagram for reversible isothermal heat transfer


We have, dQrev = T dS

Integrating between states-1 and 2,


2
Qrev = Q = ⌠
⌡ T dS
1

The integral on the RHS is equal to the area under the curve of T v/s S.
Chemical Engineering Thermodynamics - I 4.37 Second Law of Thermodynamics

For the system taken from state-1 to state-2, the area under the curve of T v/s S is equal to the
heat transferred in the reversible process, Qrev.

T Qrev

Fig. 4.21 : Heat transfer in a reversible process

The Carnot cycle consisting of two isothermal processes and two adiabatic processes is
represented by a rectangle on a T-S diagram.

1 Q1
T1 2

WC WE
T

T2
4 Q2 3

Fig. 4.22 : Carnot cycle on T-S plane

Process 1-2 : Reversible heat addition Q1 to the system at T1 from a source.

Process 2-3 : Reversible adiabatic expansion of the system producing work WE.

Process 3-4 : Reversible isothermal heat rejection from the system at T2 to a sink.

Process 4-1 : Reversible adiabatic compression of the system requiring work WC.

The area under the curve 1-2 indicates the heat Q1 added to the system [= T1 (S2 – S1)], the
area under the curve 3-4 indicates the heat Q2 rejected by the system [= T2 (S3 – S4)] and the area
bounded by points 1, 2, 3 and 4 indicates the net work output W [= Wnet = WE – WC].
Chemical Engineering Thermodynamics - I 4.38 Second Law of Thermodynamics

The efficiency of a Carnot engine is given by


W Q1 – Q2
η = Q = Q1
1

T1 (S2 – S1) – T2 (S3 – S4)


η = T1 (S2 – S1)
T1 (S2 – S1) – T2 (S2 – S1)
= T1 (S2 – S1) , since S3 = S2 and S4 = S1

T1 – T2 T2
= T1 = 1 – T1
W = Wnet = Q1 – Q2
= T1 (S2 – S1) – T2 (S3 – S4) = T1 (S2 – S1) – T2 (S2 – S1)
W = (T1 – T2) (S2 – S1) … (4.57)
It is clear from the T-S diagram that the entropy change Q1/T1 in process 1-2, where Q1 is
transferred reversibly (heat addition) at T1 is exactly equal to the entropy change – Q2/T2 in
process 3-4, where – Q2 is transferred reversibly (heat rejection) at T2. That is,
Q1 – Q2
T1 = T2
Q1 Q2
or T1 + T2 = 0
Entropy change for an ideal gas :
The first law of thermodynamics for a closed system is
dU = dQ – dW … (4.58)
For a reversible process and displacement work (work of expansion),
dW = PdV
where P is the pressure of the system.
∴ dU = dQ – PdV ... (4.59)
We know that : dQ = TdS ... (4.60)
∴ dU = TdS – PdV ... (4.61)
Suppose that an ideal gas undergoes a change in state from (T1, V1) to (T2, V2).
For an ideal gas, Cv is given by
dU
Cv = dT
∴ dU = Cv dT ... (4.62)
where Cv is the molar heat capacity at constant volume.
Rearranging Equation (4.61) gives
TdS = dU + PdV ... (4.63)
The ideal gas equation for one mole of an ideal gas is
PV = RT
RT
∴ P = V ... (4.64)
Chemical Engineering Thermodynamics - I 4.39 Second Law of Thermodynamics

Substituting for dU from Equation (4.62) and for P from Equation (4.64), Equation (4.63)
RT
becomes TdS = Cv dT + V dV
dT dV
∴ dS = Cv T + R V … (4.64 a)
Integrating between the limits, assuming that Cv is independent of temperature gives
T2 V2
S2
dT dV
⌠ dS = Cv ⌠ T + R ⌠ V

S
⌡ ⌡
1 T1 V1

∆S = (S2 – S1) = Cv ln T  + R ln V 


T2 V2
... (4.65)
 1  1
For 'n' moles of an ideal gas, Equation (4.65) becomes
T2 V2
∆S = n Cv ln T  + nR ln V  ... (4.66)
 1  1
For a constant volume process, V1 = V2 and therefore, Equation (4.66) becomes
T2
∆S = n Cv ln T  … for an isochoric process ... (4.67)
 1
For a constant temperature process, T1 = T2 and therefore, Equation (4.66) becomes

∆S = nR ln V  … for an isothermal process


V2
... (4.68)
 1
For one mole of an ideal gas, the ideal gas equation at state 1 is
P1V1 = RT1
Similarly, at state 2, P2V2 = RT2
P2V2 T2
∴ P1V1 = T1
V2 P1T2
Rearranging, we get V1 = P2T1
For an ideal gas, Cp and Cv are related by the relation
Cp – Cv = R
where R is the universal gas constant.
Cv = Cp – R
V2
Substituting values of Cv and V in Equation (4.65), we get
1

∆S = (Cp – R) ln T  + R ln P T 


T2 P1T2
 1  2 1
∆S = Cp ln T  – R ln T  + R lnT  + R ln P 
T2 T2 T2 P1
 1  1  1  2

 2
T  1
∆S = Cp ln T  + R ln P 
P
... (4.69)
 1  2
where Cp is the molar heat capacity at constant pressure.
For n moles of an ideal gas, the above equation becomes
Chemical Engineering Thermodynamics - I 4.40 Second Law of Thermodynamics

 T 2 P1
∆S = n Cp ln  T  + nR ln P  … (4.70)
 1  2
For a constant temperature process, i.e., during an isothermal expansion process, T is
constant and thus T1 = T2. Therefore, the above equation becomes

∆S = nR ln P  = nR ln V 
P1 V2
... (4.71)
 2  1
P1 V2
(Since P = V at constant T)
2 1

For a constant pressure process (i.e., for an isobaric process), P is constant and thus P1 = P2.
Therefore, Equation (4.70) becomes

∆S = n Cp ln T 
T2
... (4.72)
 1
In all the above equations,
∆S = entropy change, J/K
n = number of moles of gas, mol
T2 = final temperature of gas, K
T1 = initial temperature of gas, K
V2 = final volume of gas
V1 = initial volume of gas
P2 = final pressure of gas
P1 = initial pressure of gas
Cv = molar heat capacity of gas at constant volume,
J/(mol.K)
Cp = molar heat capacity of gas at constant pressure,
J/(mol.K)
R = universal gas constant = 8.31451 J/(mol.K)
Entropy change during Adiabatic expansion :
In such a process, Q = 0 at all stages. Hence, ∆S = 0. Thus, a reversible adiabatic process is
called as an isentropic process (process of constant entropy).
Entropy change during phase change :
A phase change operation takes place at a definite temperature and pressure and is
accompanied by absorption or evolution of heat.
Chemical Engineering Thermodynamics - I 4.41 Second Law of Thermodynamics

The entropy change associated with a phase change [e.g., from solid to liquid (fusion) or
from liquid to vapour (vaporisation)], carried out reversibly at a definite temperature
(melting/fusion point or boiling point) can be evaluated from values of the latent heat of phase
change and the temperature at which the phase change occurs.
In a phase change operation, Qrev is equal to the latent heat of phase change.
Qrev Latent heat of phase change
∆Sphase change = T = Temperature at which phase change occurs
Suppose one kg (or kmol) of a substance changes from liquid to vapour state reversibly at its
boiling point Tb at constant pressure. Let ∆Hvap be the latent heat of vaporisation, then the
entropy change accompanying the process of vaporisation is given by
∆Hvap
∆Svap = Tb … (4.73)

Similarly, if a substance changes from solid to liquid state reversibly at its melting point or
fusion point Tf at constant pressure and ∆Hfusion is the latent heat of fusion, then the entropy
change associated with this fusion process is given by
∆Hfusion
∆Sfusion = Tf … (4.74)

Since ∆Hvap (or ∆Hv) and ∆Hfusion (or ∆Hf) are both positive, both the fusion/melting and
vaporisation processes are accompanied by increase of entropy.
For a change of state from vapour to liquid and from liquid to solid, both ∆Hvap and ∆Hfusion
will be negative and therefore, the processes of condensation and freezing would be accompanied
by decrease of entropy. [∆Hvap or ∆Hfusion may be expressed in J/kg, kJ/kg, kJ/kmol or J/mol].
Third Law of Thermodynamics
The entropy of any substance at the temperature T and a given pressure P can be given by
T T
Cp
S – S0 = ⌠
⌡ T dT = ⌠
⌡ Cp d ln T … (4.75)
0 0

where S0 is the hypothetical entropy at the absolute zero of temperature. It is possible to calculate
the value of the entropy (absolute value) at any required temperature from heat capacity data,
provided the value of S0 is known.
The third law of thermodynamics states that : at the absolute zero of temperature, the
entropy of a perfect crystalline substance is zero.
According to this law, the entropies of all pure crystalline solids may be taken as zero at the
absolute zero of temperature. The law is true only for the substances which exist in the perfectly
crystalline form at 0 K (– 273 oC). If imperfections of any type are there in the perfect crystalline
arrangement at 0 K, then the entropy will be larger than zero.
Chemical Engineering Thermodynamics - I 4.42 Second Law of Thermodynamics

The third law of thermodynamics helps us to calculate the absolute entropies of pure
substances at different temperatures. The entropy S of any substance at different temperature T
may be calculated from the heat capacity data for the substance between 0 K and T K by using
the relation
T

S – S0 = ⌠
⌡ Cp d ln T
0
but S0 = 0 at T=0K
T

∴ S = ⌠
⌡ Cp d ln T … (4.76)
0
For a liquid or gaseous substance, the total absolute entropy of the substance at a given
temperature is equal to the sum of all the entropy changes involved in processes which the
substance has to undergo to reach the required state from a crystalline solid at absolute zero.
For example, for a gas at T (e.g., 298 K) and 1 bar, it may be given by
Tf Tb
∆Hfusion
S(T) = ⌠
⌡ Cp (solid) d ln T + Tf +⌠
⌡ Cp (liquid) d ln T
0 Tf
T
∆Hvap.

+ Tb + ⌡ Cp (gas) d ln T … (4.77)
Tb

(Solid at T = 0 K → Solid at Tf → Liquid at Tf → Liquid at Tb → Gas/Vapour at Tb →


Gas at T = 298 K) where Tf is the fusion/melting point, Tb is the boiling point.
∆Hfusion
Entropy of fusion = ∆Sfusion = Tf
∆Hvap.
Entropy of vaporisation = ∆Svap = Tb .
(i) If the substance is solid at temperature T, then only the first term of Equation (4.77) is
to be taken into account/is to be considered with Tf = T for the calculation of S(T).
(ii) If the substance is liquid at temperature T, then the first three terms of Equation (4.77)
are to be taken into account/are to be considered with Tb = T for the calculation of S(T).
(iii) If the substance is gas at temperature T, then all the terms of Equation (4.77) are to be
taken into account for the calculation of absolute entropy [S(T)].
Entropy change in Chemical reactions :
The entropy change of a chemical reaction is equal to the difference between the sum of
entropies of the products and the sum of entropies of the reactants. For example, for the reaction,
aA + bB + … = rR + sS + …
∆S = (rSR + sSS + …) – (aSA + bSB + …)
i.e., ∆S = ∑ Sproducts – ∑ Sreactants … (4.78)
Chemical Engineering Thermodynamics - I 4.43 Second Law of Thermodynamics

The entropy of 1 mol of a substance in the pure state at 1 atm pressure and 298 K (25°C) is
called the standard entropy of the substance. It is denoted by the symbol So. Hence,
o o
∆So = ∑ Sproducts – ∑ Sreactants … (4.79)

Entropy generation in a closed system :


2
dQ dQ
We know that, dS ≥ T or ∆S ≥ ⌠
⌡ T … (4.80)
1

The equality sign holds good for a reversible process and the inequality sign holds good for
an irreversible process.
For any process, the above equation can be written as
dQ
dS = T + dSG
2
dQ
or ∆S = S2 – S1 = ⌠
⌡ T + SG … (4.81)
1

2
dQ
where SG denotes the entropy generated or produced due to internal irreversibility and ⌠
⌡ T
1
represents entropy transfer along with heat.
2
dQ
If the process is reversible, SG = 0 and therefore ∆S = ⌠
⌡ T .
1

2
dQ
SG = ∆S – ⌠
⌡ T
1

2
dQ
Since, ∆S ≥ ⌠
⌡ T
1

SG ≥ 0
2
dQ
If the process is reversible, ∆S = ⌠
⌡ T and therefore SG = 0.
1

2
dQ
If the process is irreversible, ∆S > ⌠
⌡ T and therefore SG > 0.
1

Hence, SG may be positive or zero but it cannot be a negative quantity.


For an isolated system, dQ = 0. Therefore, from Equation (4.81),
∆S = SG ≥ 0 … (4.82)
Chemical Engineering Thermodynamics - I 4.44 Second Law of Thermodynamics

Entropy generation in an open system :


In an open system, mass, energy and entropy cross the boundary of the system.
For a steady-state flow process, an entropy balance gives
·
· · Q
SG = m (Se – Si) – T … (4.83)
·
where SG is the rate of entropy generation due to irreversibilities within the system in W/K,
· · ·
Q is the rate of heat transfer, Q /T is the rate of entropy transfer along with heat, m is the mass
flow rate of the flowing fluid, Se is the entropy of the fluid at the exit/outlet and Si is the entropy
of the fluid at the inlet.
·
If the flow device is adiabatic, Q = 0. Therefore, for an adiabatic flow device operated in a
steady-state fashion, Equation (4.83) reduces to
· ·
SG = m (Se – Si) ≥ 0 … (4.84)
It is clear from the above equation that if the flow is steady and adiabatic, the entropy is
generated and not conserved in any real flow process.
If the flow process is reversible, then Equation (4.84) yields
Se – Si = 0
or Se = Si … (4.85)
Hence, during a steady, reversible and adiabatic flow through a device, the entropy of the
flowing fluid remains constant.
Isentropic efficiency :
A reversible and adiabatic device is called an isentropic device.
The performance of devices such as turbines (work producing) and compressors and pumps
(work consuming) is generally expressed in terms of isentropic efficiency.
(a) Turbines : A turbine is a work/power producing device. Therefore, for the same inlet
conditions and exit pressure,
(power output of an isentropic turbine) > (power output of an actual turbine)
The isentropic efficiency of a turbine is given by
Power output of the actual turbine per unit mass of fluid
ηT = Power output of the same turbine per unit mass of fluid‚ if it were isentropic

(b) Compressor/Pumps : A compressor (or pump) is a power consuming device. Therefore,


for the same inlet conditions and exit pressure,
Power consumption of an isentropic < Power consumption of an
 compressor/pumps   actual compressor/pump 
The isentropic efficiency of a compressor/pump is given by
Power required by the compressor/pump per unit mass of fluid‚
if it were isentropic
ηC or P = Power required by the actual compressor/pump per unit mass of fluid
Chemical Engineering Thermodynamics - I 4.45 Second Law of Thermodynamics

Combined First and Second laws :


For a closed system, by the first law,
dU = dQ – dW … in differential form
dQ = dU + dW
By the second law, dQ = dQrev = T dS
For a reversible process,
dW = P dV … displacement work
Therefore, substituting for dW, we get
dQ = dU + P dV
Substituting for dQ, we get
T dS = dU + P dV … (4.86)
The enthalpy is given by
H = U + PV
The differential enthalpy is given by
dH = dU + P dV + V dP
Substituting for dU + P dV, the above equation becomes
dH = T dS + V dP
i.e., T dS = dH – V dP … (4.87)
Equations (4.86) and (4.87) are known as fundamental thermodynamic relations combining
the first and second laws of thermodynamics into a single equation.
Eventhough Equations (4.86) and (4.87) are derived for a reversible process, they are also
applicable for any irreversible process since they express relationships among the properties of
the closed system, which are independent of the path.
The actual work done by a system is always less than the idealised reversible work, i.e., the
work done by a system when the process is irreversible is less than that when the process is
reversible. Therefore, the difference (P dV – dW) indicates the work that is lost due to
irreversibility.
Let us examine the equations obtained from the first and second laws of thermodynamics :
(1) dE = dQ – dW, i.e., dQ = dE + dW
This equation is applicable for any system and for any process, reversible or
irreversible.
(2) dU = dQ – dW, i.e., dQ = dU + dW
This equation is applicable for any process, reversible or irreversible and for a closed
(non-flow) system.
(3) dU = dQ – P dV, i.e., dQ = dU + P dV
This equation is applicable for a closed system undergoing a reversible process
involving displacement work.
Chemical Engineering Thermodynamics - I 4.46 Second Law of Thermodynamics

(4) dQ = T dS
This equation is applicable only for a reversible process.
(5) T dS = dU + P dV
This equation is applicable for any process reversible or irreversible undergone by a
closed system since this equation expresses a relationship among the thermodynamic
properties of the system, which are independent of the path.
(6) T dS = dH – V dP
This equation is applicable for any process reversible or irreversible, undergone by a
closed system since it is a relationship among the thermodynamic properties of the
system, which are independent of the path followed.
Comparison of First law of thermodynamics with the Second law of thermodynamics :
1. According to the first law, all forms of energy are equivalent (work and heat are of
equal quality).
1. According to the second law, all forms of energy are not equivalent qualitywise (work is
superior to heat).
2. This law states that in any cyclic process either heat can be completely converted into
work or work can be completely converted into heat.
2. This law states that heat from a constant temperature source cannot be completely
converted into work in any cyclic process.
3. It may be stated to be the law of internal energy.
3. It may be stated to be the law of entropy.
4. According to this law, a PMM-1 is not possible.
4. According to this law, a PMM-2 is not possible.
5. It does not give us an idea regarding the extent of a process.
5. It gives us an idea regarding the extent of a process.
6. It does not give us an idea regarding the feasibility of a process. Moreover according to
this law all processes are possible provided that energy is conserved.
6. It gives us an idea regarding the feasibility of a process.
7. In every real process, energy is conserved (quantity wise).
7. In every real process, energy is conserved, but its quality is degraded (i.e., energy
becomes less available for the performance of work.)
8. It states that the total energy of an isolated system can neither be created nor destroyed
(i.e., it remains constant).
8. It states that the entropy of an isolated system (in all real processes) cannot be destroyed
but it can be created (i.e., it increases).
9. It is a necessary but not a sufficient condition for a real process to take place.
9. Along with the first law it forms the necessary and sufficient condition for a real
process to occur/take place.
Chemical Engineering Thermodynamics - I 4.47 Second Law of Thermodynamics

SOLVED EXAMPLES
Example 4.1 :
10 kg of water (the system) is heated from 290 K to 340 K by
(i) using saturated steam at 10 bar,
(ii) using superheated steam at 5 bar and 523 K.
Assume that the condensate is not subcooled in both the cases.
Calculate the change in entropy of the system, the entropy change of the surroundings and
the total entropy change in each case.
Data : Saturated steam at 10 bar : H = 2778.10 kJ/kg,
H of condensate = 762.81 kJ/kg, T of condensate = 453 K
S of steam = 6.5865 kJ/(kg·K), S of condensate = 2.1387 kJ/(kg·K)
Superheated steam at 5 bar and 523 K : H = 2960.7 kJ/kg, S = 7.2709 kJ/(kg·K)
H of condensate = 640.23 kJ/kg, T of condensate = 425 K,
S of condensate = 1.8607 kJ/(kg·K)
Solution :
In this problem,
System = Water
Surroundings = Steam
System : Water : m = 10 kg, Cp = 4.187 kJ/(kg·K), T1 = 290 K and T2 = 340 K
Heating of water is a constant pressure process.
(i) Case - I : Use of saturated steam at 10 bar
The change in entropy of the system (water) is

∆Ssystem = mCp ln T 


T2
 1
= 10 × 4.187 ln 290 = 6.66 kJ/K
340
... Ans. (i)
 
The change in entropy of steam is the difference between the entropy of the condensate and
saturated steam.
Change in entropy of steam = S of condensate – S of saturated steam
= 2.1387 – 6.5865
= – 4.4478 kJ/(kg·K)
The enthalpy balance gives the mass of steam used.
Heat given out by steam = Heat gained by water
m (2778.1 – 762.81) = 10 × 4.187 × (340 – 290)
∴ m = 1.0388 ≈ 1.039 kg
Chemical Engineering Thermodynamics - I 4.48 Second Law of Thermodynamics

The change in entropy of the surroundings (steam) is


∆Ssurroundings = [– 4.4478 kJ/(kg·K)] × 1.039
= – 4.621 kJ/K ... Ans.
The total entropy change is
∆Stotal = ∆Ssystem + ∆Ssurroundings
= 6.66 + (– 4.621) = 2.039 kJ/K ... Ans. (i)
(ii) Case - II : Use of superheated steam at 5 bar and 523 K.
The entropy change of the system (water) is
T2
∆Swater = mCp ln T 
 1
= 10 × 4.187 ln 290 = 6.66 kJ/K
340
... Ans. (ii)
 
The mass of superheated steam required can be obtained by an enthalpy balance.
Heat given out by steam = Heat gained by water
m (H of superheated steam – H of condensate) = mCp (T2 – T1)
m (2960.7 – 640.23) = 10 × 4.187 (340 – 290)
∴ m = 0.9022 kg
The change in entropy of the surroundings (i.e., steam) is the difference between the entropy
of the condensate and the entropy of the superheated steam.
∆Ssurroundings = m of superheated steam [S of condensate – S of superheated steam]
= 0.9022 (1.8607 – 7.2709)
= – 4.88 kJ/K ... Ans. (ii)
The total entropy change is
∆Stotal = ∆Ssystem + ∆Ssurroundings
= 6.66 + (– 4.88) = 1.78 kJ/K ... Ans. (ii)
Example 4.2 :
One mole of an ideal gas is compressed isothermally at 400 K from 100 kPa to 1000 kPa.
The work required for this irreversible process is 20% more than that for a reversible
compression. The heat liberated during the process of compression is absorbed by a thermal
reservoir at 300 K.
Calculate : (i) the entropy change of the gas, (ii) the entropy change of the reservoir and
(iii) the total entropy change.
Solution :
The gas is to be compressed from P1 = 100 kPa to P2 = 1000 kPa at T = 400 K
For an isothermal process, we have

∆S = R ln P 
P1
 2
Chemical Engineering Thermodynamics - I 4.49 Second Law of Thermodynamics

The entropy change of the gas is

∆Sgas = R ln P 
P1
 2
= 8.31451 ln 1000
100
 
= – 19.14 kJ/(kmol·K) ... Ans.
Work required for a reversible compression at constant T is given by

W = RT ln P 
P1
 2
= 8.31451 × 400 × ln 1000
100
 
= – 7657.94 ≈ 7658 kJ per mol
The minus (–ve) sign indicates that the work is done on the system.
Work required for the actual compression process (irreversible compression) is 20% more
than that for the reversible compression of the gas.
Wirr = 1.20 × (– 7658)
= – 9189.6 kJ per mol
For an isothermal process :
Q = W
∴ Q = – 9189.6 kJ per mol
where the minus sign signifies that heat is liberated.
This heat will be absorbed by the reservoir at 300 K.
Heat absorbed by the reservoir = Q = 9189.6 kJ
The entropy change of the reservoir is
Q 9189.6
∆Sreservoir = T = 300 = 30.632 kJ/K ... Ans.
The total entropy change is
∆Stotal = ∆Sgas + ∆Sreservoir
= – 19.14 + 30.632 = 11.492 kJ/K ... Ans.
Example 4.3 :
A small metallic object 4 kg in mass at a temperature of 500 K is thrown into the lake which
is at 300 K. Calculate the change in entropy of the universe.
Data : Cp of object = 0.50 kJ/(kg·K)
Solution :
Consider the metallic object as the system and the lake as the surroundings. The universe is
metallic object + lake.
∆Suniverse = ∆Sobject + ∆Slake
Chemical Engineering Thermodynamics - I 4.50 Second Law of Thermodynamics

The metallic object at T1 = 500 K will lose heat Q, so its entropy will decrease. This process
of rejection of heat will occur at constant pressure.
The change in entropy of the metallic object is

∆Sobject = mCp ln T  where T1 = 500 K and T2 = 300 K


T2
 1

= 4 × 0.50 × ln 500
300
 
= – 1.02 kJ/K
The lake is at T = 300 K. It will receive heat from the metallic object.
Heat received by lake = Heat given out by metallic object
Q = mCp (T2 – T1)
= 4 × 0.50 × (500 – 300) = 400 kJ
The entropy change of the lake is
Q
∆Slake = ∆Ssurroundings = T

400
= 300 = 1.33 kJ/K

The change in entropy of the universe is


∆Suniverse = ∆Ssystem + ∆Slake
= – 1.02 + 1.33 = 0.31 kJ/K ... Ans.
Example 4.4 :
A block of 10 kg ice at 273 K is dumped in an insulated vessel that contains 100 kg water at
303 K. Calculate the entropy change of the mixture.
Data : Cp of water = 4.23 kJ/(kg·K) and latent heat of melting of ice at 273 K = 333.5 kJ/kg
Solution :
Ice : m = 10 kg, T = 273 K, Latent heat of melting = 333.5 kJ/kg
Water : m' = 100 kg, T = 303 K, Cp = 4.23 kJ/(kg·K)
Let T' be the final temperature attained by the mixture.
Then the enthalpy balance gives
Heat gained by ice = Heat lost by water
m × latent heat + mCp ∆T = m'Cp ∆T (sensible heat)
10 × 333.5 + 10 × 4.23 (T' – 273) = 100 × 4.23 (303 – T')
∴ T’ = 293.1 K
Ice at 273 K gets converted to water at 273 K and then the temperature of this water is raised
from 273 K to 293.1 K.
Chemical Engineering Thermodynamics - I 4.51 Second Law of Thermodynamics

The change in entropy of ice is

∆Sice = m 
Latent heat of melting  T' 
 + mCp ln T 
 T m   m
where Tm = melting point

= 10 × 273 + 10 × 4.23 ln  273 


333.5 293.1
 
= 15.22 kJ/K
The change in entropy of water is
T'
∆Swater = m'Cp ln T where T = 303 K, T' = 293.1 K
 

= 100 × 4.23 ln  303 


293.1
 
= – 14.05 kJ/K
The total entropy change or the entropy change of the mixture is
∆S = ∆Sice + ∆Swater
= 15.22 + (– 14.05)
= 1.17 kJ/K ... Ans.
[Entropy generated = 1.17 kJ/K]
Example 4.5 :
35 kg of a steel casting at 725 K is quenched in 150 kg oil at 275 K. If there are no heat
losses, calculate the change in entropy.
Data : Cp of steel = 0.88 kJ/(kg·K) and Cp of oil = 2.5 kJ/(kg·K)
Solution :
Steel casting : 35 kg at T = 725 K, Cp = 0.88 kJ/(kg·K)
Oil : 150 kg at T = 275 K, Cp = 2.5 kJ/(kg·K)

Let T' be the final temperature attained by the system (casting + oil).
The heat balance is
Heat lost by casting = Heat gained by oil ... Sensible heat transfer
(mCp)casting (725 – T') = (mCp)oil (T' – 275)

35 × 0.88 (725 – T') = 150 × 2.5 (T' – 275)


∴ T' = 309.15 K
Let ∆S1 be the change in entropy of the casting and ∆S2 be the change in entropy of the oil.
Chemical Engineering Thermodynamics - I 4.52 Second Law of Thermodynamics

The change in entropy of the casting is

∆S1 = mCp ln  T 
T'
 
where m = 35 kg, T' = 309.15 K and T = 725 K

∆S1 = 35 × 0.88 ln  725 


309.15
 
= – 26.25 kJ/K
The change in entropy of the oil is

∆S2 = mCp ln  T 
T'
 
where m = 150 kg, T' = 309.15 K and T = 275 K

∆S2 = 150 × 2.5 ln  275 


309.15
 
= 43.89 kJ/K
The change in entropy of the casting and oil together is
∆S = ∆S1 + ∆S2
= – 26.25 + 43.89 = 17.64 kJ/K ... Ans.
Example 4.6 :
m1 kg of a liquid at temperature T1 is mixed with m2 kg of the same liquid at temperature
T2 (T2 > T1). If the heat capacity of the liquid is Cp, show that the entropy change of the universe
due to the mixing process for equal masses (m1 = m2 = m) of the liquid is given by
(T1 + T2)/2
∆S = 2mCp ln 
 T1T2 
Also show that ∆S is always positive. Assume adiabatic mixing.
Solution :
Let Tf be the final temperature attained after mixing. Tf can be determined by an enthalpy
balance.
Heat rejected/lost by hot liquid at T2 = Heat gained by cold liquid at T1
∴ m2Cp (T2 – Tf) = m1Cp (Tf – T1)
m1T1 + m2T2
∴ Tf = m1 + m2
This mixing process takes place at constant pressure. The change in entropy during a
constant pressure process is given by

∆S = mCp ln T 
Tfinal
 initial
Chemical Engineering Thermodynamics - I 4.53 Second Law of Thermodynamics

Let ∆S1 be the entropy change of liquid at T1 and ∆S2 be that of liquid at T2.
 Tf 
∆S1 = m1Cp ln T 
 1
 Tf 
∆S2 = m2Cp ln T 
 2
Substituting for Tf in the above equation gives,
 Tf 
∆S1 = m1Cp ln T  = m1Cp ln (m + m ) T 
m1T1 + m2T2
 1  1 2 1

 Tf 
∆S2 = m2Cp ln T  = m2Cp ln (m + m ) T 
m1T1 + m2T2
 1  1 2 2

∆Ssystem = ∆S1 + ∆S2


m1T1 + m2T2 m1T1 + m2T2
= m1Cp ln (m + m ) T  + m2Cp ln (m + m ) T 
 1 2 1  1 2 2

The mixing process is adiabatic, so Q = 0.


Since no heat exchange with the surroundings, it will remain unaffected. Therefore, the
entropy of the surroundings will not change.
∴ ∆Ssurroundings = 0
The entropy change of the universe is
∆Suniverse = ∆Ssystem + ∆Ssurroundings

= m1Cp ln (m + m ) T  + m2Cp ln (m + m ) T  + 0


m1T1 + m2T2 m1T1 + m2T2
 1 2 1  1 2 2

= m1Cp ln (m + m ) T  + m2Cp ln (m + m ) T 


m1T1 + m2T2 m1T1 + m2T2
 1 2 1  1 2 2

For equal masses of the liquid at T1 and that at T2,


m1 = m2 = m
The above relation becomes,
m (T1 + T2) + mC ln m (T1 + T2)
∆Suniverse = mCp ln  2mT   2mT 
 p
1   2 

= mCp ln 
(T1 + T2)/2 (T1 + T2)/2
 + mCp ln  
T 1  T  2 
(T1 + T2)/2 (T1 + T2)/2 
= mCp ln   + ln  
  T1   T2 

 (T1 + T2)/2 × (T1 + T2)/2 


= mCp ln  
  T 1T 2  
[(T1 + T2)/2]2 
= mCp ln  
  T 1T 2 
Chemical Engineering Thermodynamics - I 4.54 Second Law of Thermodynamics

 (T1 + T2)/22 
= mCp ln 
T1T2  

 

 (T1 + T2)/2 
∆Suniverse = 2mCp ln 
T1T2  
 ... Ans.
 

 (T1 + T2)/2 
∆Suniverse = 2mCp ln 
T1T2  
We have, 
 
= 2mCp ln (A/B)
where A = (T1 + T2)/2 and B = T 1T 2
For any combination of the positive values of T1 and T2, A is always greater than B, A/B is
always greater than 1 and ln (A/B > 1) is always positive, so the entropy change of the universe,
∆Suniverse is always positive. ... Ans.
Example 4.7 :
An ideal gas at 300 K and 1000 kPa enters a rigid and insulated apparatus. This gas leaves
the apparatus in two streams in equal quantities, one is at 360 K and 100 kPa and the other is at
240 K and 100 kPa. Calculate the total entropy change. Is the process thermodynamically
possible ? Take Cp = 30 kJ/(kmol·K).
Solution :
Let us assume that 2 kmol of the gas at 300 K and 1000 kPa enters the apparatus. As this gas
leaves the apparatus in two streams equal in quantities, 1 kmol of the gas changes its state from
300 K and 1000 kPa to 360 K and 100 kPa and 1 kmol of the gas changes its state from 360 K
and 1000 kPa to 240 K and 100 kPa.
Stream 1 : 1 kmol gas : P1 = 1000 kPa, T1 = 300 K, P2 = 100 kPa and T2 = 360 K
Stream 2 : 1 kmol gas : P1 = 1000 kPa, T1 = 300 K, P2 = 100 kPa and T2 = 240 K
The apparatus is insulated, so Q = 0 (no heat interaction with the surroundings) and the
surroundings will remain unaffected. Therefore, ∆Ssurroundings = 0.
The total entropy change is
∆Stotal = ∆Ssystem + ∆Ssurroundings
= ∆Ssystem + 0 = ∆Ssystem

= Cp ln T  – R ln P  + Cp ln T  – R ln P 


T2 P2 T2 P2
  1  1stream 1   1  1stream 2

= 30 ln 300 – 8.31451 ln 1000 + 30 ln 300 – 8.31451 ln 1000


360 100 240 100
       
= 5.47 – (– 19.14) + (– 6.69) – (– 19.14)
= 37.06 kJ/(2 kmol of gas·K) ... one kmol each of streams 1 and 2
∴ ∆Stotal = 18.53 kJ/(kmol·K) ... Ans.
As ∆Stotal is +ve, the process is thermodynamically possible. ... Ans.
Chemical Engineering Thermodynamics - I 4.55 Second Law of Thermodynamics

Example 4.8 :
Two perfectly insulated tanks each of 1 m3 volume are connected by means of a pipeline
containing valve. Initially, one tank contains an ideal gas at 290 K and 200 kPa, while the other
is completely evacuated. The valve is opened to allow the pressures and temperatures to
equalise.
(i) Determine the final temperature and pressure in the tanks.
(ii) Determine the entropy change of the gas.
Solution :
The tanks are insulated, so Q = 0
No work is done on or by the system, so W = 0 … free expansion
Hence, ∆U = 0
The first law for a closed system is
∆U = Q – W = 0 – 0 = 0
We know that the internal energy of an ideal gas depends only on temperature, so ∆T = 0 as
∆U = 0. Therefore, initial temperature (T1) = final temperature (T2).
∴ The final temperature of both the tanks is 290 K ... Ans. (i)
Volume of each tank =1 m3
Pressure of tank-1 =200 kPa ... Tank containing the ideal gas
Final volume occupied by the gas =2 × 1 = 2 m … 2 tanks, each of V = 1 m3
3

Final temperature =Initial temperature = 290 K


P1V1 P2V2
∴ T1 = T
2

P1V1 T2
Final pressure in both tanks = P2 = T × V
1 2

200 × 1 290
= 290 × 2
= 100 kPa
∴ The final pressure is 100 kPa. ... Ans. (i)
Here the system changes from 290 K and 200 kPa to 290 K and 100 kPa.
For a reversible isothermal expansion, we have
P1
∆S = R ln P 
 2
= 8.31451 ln 100
200
 
= 5.763 kJ/(kmol·K)
Chemical Engineering Thermodynamics - I 4.56 Second Law of Thermodynamics

The amount of gas in the system is


P1V1 200 × 1
n = RT = = 0.08294 ≈ 0.0823 kmol
1 8.31451 × 290
The entropy change for the gas is
∆S = n∆S
= 0.083 × (5.763)
= 0.478 ≈ 0.48 kJ/K ... Ans. (ii)
Example 4.9 :
1 kmol of water is heated at constant pressure of 14 bar from a temperature of 294 K (21oC)
to the boiling point and then completely vaporised at this pressure. Determine the portion of the
heat transferred that is unavailable for transformation to work in a heat engine assuming the
heat sink to be at 283 K.
Data : Boiling point of water at 14 bar = 468 K
Latent heat of vaporisation = 1960 kJ/kg
Cp of water = 4.187 kJ/(kg·K)
Solution :
Amount of water to be heated and vaporised at constant P = 1 kmol = 18 kg
Initial temperature of water = T1 = 294 K
Boiling point of water at 14 bar = Tb = 468 K = T2
Latent heat of vaporisation (λ) = 1960 kJ/kg
Cp of water = 4.187 kJ/(kg·K)
Heat transferred to water is

Q1 =  water from T to T  + vaporise water


Sensible heat to heat Latent heat to
 1 b
= mCp (Tb – T1) + mλ
= 18 × 4.187 (468 – 294) + 18 × 1960
= 48393.684 ≈ 48393.7 kJ
The change in entropy for this process of heating water to vaporise it completely can be
calculated as
∆S = ∆S1 + ∆S2
∆S1 = Entropy change for heating water from T1 = 294 K to T2 = Tb = 468 K
∆S2 = Entropy change during vaporisation of water

∆S1 = mCp ln T 


T2
 1
= 18 × 4.187 ln 294 = 35.04 kJ/K
468
 
Chemical Engineering Thermodynamics - I 4.57 Second Law of Thermodynamics

m·∆Hv m × Latent heat of vaporisation


∆S2 = Tb = Boiling point temperature
18 × 1960
= 468 = 75.3846 ≈ 75.38 kJ/K
∴ Total entropy change = ∆S = 35.04 + 75.38 = 110.42 kJ/K
Heat rejected to the sink by the engine = Q2 = To ∆S
where To = Sink temperature = 283 K
Q2 = Heat that is unavailable for doing work = To ∆S = 283 × 110.42 = 31248.86 kJ
Fraction/portion of the heat transferred that is unavailable for transformation into
Q2 31248.86
work = Q = 48393.7 = 0.6457 ≈ 0.646 or 64.6% ... Ans.
1

Example 4.10 :
In a heat exchanger, cold air is heated from 293 K to 353 K by means of hot air which enters
the exchanger at 423 K. The molar flow rates of both the streams are equal. The specific heat of
air is 29.3 kJ/(kmol·K).
Calculate the entropy change of both the air streams and the total entropy change.
Solution :
Cold air : T1 = 293 K and T2 = 353 K
Hot air : T' = 423 K and T' = ?
1 2
Cp of air (both cold and hot) = 29.3 kJ/(kmol·K)
.
Molar flow rate of cold air = Molar flow rate of hot air = n
The exit temperature of the hot air (T' ) can be obtained by enthalpy balance.
2

Heat gained by cold air = Heat lost by hot air


. .
n × 29.3 (353 – 293) = n × 29.3 (423 – T' ) 2

∴ T'2 = 363 K
It is assumed that the exchanger is insulated.
∆Ssurroundings = 0 ... as the exchanger is insulated there is no heat interaction with the
surroundings and hence, the entropy of the surroundings does not change.
Let ∆S1 be the change in entropy of the cold air and ∆S2 be the change in entropy of the hot
air.
The entropy change of the cold air is

∆S1 = Cp ln T 
T2
 1
= 29.3 ln 293
353
 
= 5.458 ≈ 5.46 kJ/(kmol·K) ... Ans.
Chemical Engineering Thermodynamics - I 4.58 Second Law of Thermodynamics

The entropy change of the hot air is


T' 
∆S2 = Cp ln  ' 
2
T 
 1

= 29.3 ln 423
363
 
= – 4.48 kJ/(kmol·K) ... Ans.
The total entropy change is
∆S = ∆S1 + ∆S2 + ∆Ssurroundings
= 5.46 + (– 4.48) + 0
= 0.98 kJ/(kmol.K) ... Ans.
Example 4.11 :
4 kg of water at 300 K are mixed with 1 kg of ice at 273 K. Assume the mixing process to be
adiabatic. Determine the final temperature of mixture of ice and water and the net change of
entropy.
Enthalpy fusion of ice = 335 kJ/kg and Cp of water = 4.187 kJ/(kg·K)
Solution :
Let the final temperature of the mixture of ice and water be Tf K.
Amount of water = m = 4 kg and amount of ice = m' = 1 kg
The temperature of water will change from 300 K to Tf K.
[No phase change for water ... Only sensible heat transfer]
Ice will melt at 273 K to water at 273 K and then temperature of this water will change from
273 K to Tf K.
[Latent heat + Sensible heat transfer]
The enthalpy balance for the process is
Heat given out by water = Heat gained by ice
mCp ∆T = m'λ + m'Cp ∆T
4 × 4.187 (300 – Tf) = 1 × 335 + 1 × 4.187 (Tf – 273)
∴ Tf = 278.6 K (5.6oC)
Entropy change of water = ∆S1
Entropy change of ice = ∆S2
Entropy change of surroundings = ∆Ssurr = 0 ... Adiabatic mixing - no heat transfer to the
surroundings.
The net entropy change/Total entropy change,
∆S = ∆Ssystem + ∆Ssurroundings = ∆S1 + ∆S2 + 0 = ∆S1 + ∆S2
Chemical Engineering Thermodynamics - I 4.59 Second Law of Thermodynamics

(i) ∆S1 : For water from T1 = 300 K to T2 = 278.6 K :


T2
∆S1 = mCp ln T  where m = 4 kg
 1

= 4 × 4.187 ln  300 
278.6
 
= – 1.2394 ≈ – 1.24 kJ/K
(ii) ∆S2 : For ice from T1 = 273 K to water at T2 = 278.6 K :
Enthalpy of fusion = Latent heat of fusion = 335 kJ/kg
T1 = 273 K = Temperature at which phase change occurs (from water to ice)

∆S2 =
m' × Latent heat of fusion T2
+ m'Cp ln T 
T1  1
1 × 335 278.6
= 273 + 1 × 4.187 × ln  273 
= 1.312 kJ/K
∆S = ∆S1 + ∆S2
= – 1.24 + 1.312 = 0.072 kJ/K ... Ans.
Example 4.12 :
In an insulated heat exchanger, water (40 kg/min) is heated from 323 K (50°C) to
388 K (115oC) by hot gases (80 kg/min) that enter the heat exchanger at 493 K (220°C). Find the
entropy change of the universe.
Take : Cp for water = 4.187 kJ/(kg·K) and Cp for gases = 1.10 kJ/(kg.K)
Solution :

Let T'2 be the final temperature of gases leaving the heat exchanger. This can be determined
by an enthalpy balance.
Heat gained by water = Heat removed from hot gases
. .
mwCpw (T2 – T1) = mgCpg T1' – T2'
( )
.
where T1 = 323 K, T2 = 388 K, Cpw = 4.187 kJ/(kg·K), mw = 40 kg/min,
.
mg = 80 kg/min, Cpg = 1.10 kJ/(kg·K), T2' = ? and T1' = 493 K

∴ 40 × 4.187 (388 – 323) = 80 × 1.10 × 493 – T2' ( )


∴ T'2 = 369.3 K
Chemical Engineering Thermodynamics - I 4.60 Second Law of Thermodynamics

Entropy change = Entropy change of the water + Entropy change of the hot
 of the universe 
gases + Entropy change of the surroundings
∆Suniverse = ∆S1 + ∆S2 + ∆Ssurroundings
[∆S1 + ∆S2 = ∆Ssystem]
Entropy of the surroundings does not change as the surroundings are unaffected ... the heat
exchanger is insulated … no heat interaction with the
surroundings (Q = 0).
∴ ∆Ssurroundings = 0
∆Suniverse = ∆S1 + ∆S2 + 0 = ∆S1 + ∆S2
T' 
. T 2
= mw Cpw ln T  + mg Cpg ln  ' 
. 2
 1 T 
 1

= 40 × 4.187 × ln 323 + 80 × 1.1 × ln  493 


388 369.3
   
= 30.708 – 25.423 = 5.285 kJ/K ... per min basis ... Ans.
Example 4.13 :
Calculate the change in entropy when 2 mol of ice are heated from 263 K to 283 K.
Data : Cp of ice = 37.7 J/(mol·K), Cp of water = 75.3 J/(mol·K)
and molar enthalpy of fusion = 6.01 kJ/mol at 273 K
Solution :
The total entropy change is
∆S = ∆S1 + ∆S2 + ∆S3
∆S1 = Entropy change when 2 mol of ice are heated from
T1 = 263 K to T2 = 273 K (0 °C)

∆S1 = nCp (ice) ln T 


T2
 1
where n = 2 mol, Cp = 37.7 J/(mol·K), T1 = 263 K and T2 = 273 K

∆S1 = 2 × 37.7 × ln 263 = 2.814 J/K


273
 
∆S2 = Entropy change during melting of ice at 273 K (0oC)
n · ∆Hf
= Tf
where Tf = 273 K and ∆Hf = 6.01 kJ/mol = 6010 J/mol
2 × 6010
∴ ∆S2 = 273 = 44.03 J/mol
Chemical Engineering Thermodynamics - I 4.61 Second Law of Thermodynamics

∆S3 = Entropy change for heating of water from T1 = 273 K


to T2 = 283 K
T2
∆S3 = n·Cp ln T 
 1
where Cp = 75.3 J/(mol·K)

∆S3 = 2 × 75.3 × ln 273 = 5.42 J/K


283

 
Total entropy change for converting ice at 263 K to water at 283 K is
∆S = 2.814 + 44.03 + 5.42
= 52.264 J/K ... Ans.
Example 4.14 :
A rigid and insulated tank of 1 m3 volume containing O2 at 500 K and 1 MPa is connected to
a second insulated tank through a valve. The second tank is of 2 m3 volume and contains N2 at
800 K and 2 MPa. The valve is opened and gases are allowed to mix and reach an equilibrium
state. Determine the change in entropy.
Assume ideal gas behaviour of O2 and N2 with γ = 1.4.
Solution :
(i) O2 in tank - 1 :
V = 1 m3, T = 500 K and P = 1 MPa = 1000 kPa
PV
nO2 = moles of O2 = RT

1000 × 1
= = 0.240 kmol = 240 mol
8.31451 × 500
(ii) N2 in tank - 2 :
V = 2 m3, T = 800 K and P = 2 MPa = 2000 kPa
2000 × 2
nN2 = moles of N2 =
800 × 8.31451
= 0.601 kmol = 601 mol
Consider the gases contained in both the tanks as the system.
Both the tanks are insulated, so Q = 0
No work is done on or by the system (gases), so W = 0 … simply mixing of gases
The system (the gases contained in the tanks) exchanges no heat and work with the
surroundings. Then, by the first law of thermodynamics,
∆U = Q – W = 0 (since Q = 0 and W = 0)
∴ U1 = U2
i.e., the internal energy after mixing of gases remains the same as that before mixing.
Chemical Engineering Thermodynamics - I 4.62 Second Law of Thermodynamics

Let T be the final temperature attained after mixing.


We have : U1 = U2
(nCv)O TO2 + (nCv)N TN2 = [(nCv) O2
+ (nCv)N ] T
2 2 2

Cv of O2 = Cv of N2 as γ = 1.4 for N2 and O2


nO 2 T O 2 + nN 2 T N 2 = [ nO 2 + nN 2 ] T
240 × 500 + 601 × 800
∴ T = 240 + 601
= 714.38 ≈ 714.4 K
After mixing :
Final V = V1 + V2 = 1 + 2 = 3 m3
Final T = 714.4 K
Final moles = moles of O2 + moles of N2
= 240 + 601 = 841 mol (= n)
nRT
Final pressure, P = V

841 × 8.31451 × 714.4


= 3
= 1665148 Pa [as R = 8.31451 m3·Pa/(mol·K)]
= 1665.148 kPa = 1.665 MPa
Hence, 240 mol O2 gas changes its state from 500 K (T1) and 1 MPa (P1) to 714.4 K (T2) and
1.665 MPa (P2), while 601 mol N2 gas changes its state from 800 K (T1) and 2 MPa (P1) to
714.4 K (T2) and 1.665 MPa (P2).
Given : Ideal gas behaviour of O2 and N2 with γ = 1.4.
Cp and Cv will be the same for both the gases.
Cp
Cv = γ = 1.4 and Cp – Cv = R
Cp = 1.4 Cv
∴ 1.4 Cv – Cv = R
∴ 0.4 Cv = R
∴ Cv = 2.5 R
Cp – 2.5 R = R
∴ Cp = 3.5 R = 3.5 × 8.31451 kJ/(kmol.K)
= 29.10 kJ/(kmol·K)
= 29.10 J/(mol·K)
Chemical Engineering Thermodynamics - I 4.63 Second Law of Thermodynamics

The entropy change per mole of an ideal gas is given by

∆S = Cp ln T  – R ln P 
T2 P2
... when both T and P are changing
 1  1
In our case, ∆S = ∆S for O2 + ∆S for N2
 T2 P2  T2 P2
∆S = nO2 Cp ln T  – R ln P  + nN2 Cp ln T  – R ln P 
  1  1   1  1
= 240 29.10 ln  500  – 8.31451 ln  1 
714.4 1.665
    
+ 601 29.10 ln  800  – 8.31451 ln  1 
714.4 1.665
    
= 1474.79 + (– 1063.15)
= 411.64 J/K ... Ans.
∆Stotal = ∆Ssystem + ∆Ssurroundings
[This is the total entropy change as ∆Ssurroundings = 0 ... as the tanks are insulated, no heat
interaction, Q = 0]
Example 4.15 :
Liquid helium is produced from its saturated vapour at 4.2 K. The latent heat of vaporisation
of helium at this temperature is 83.3 kJ/kmol. Determine the minimum work required to produce
liquid helium if the temperature of the ambient atmosphere is 305 K.
Solution :
The minimum work required is given by
Q (T1 – T2)
W = T2
where T1 = Ambient temperature = 305
T2 = Tb = 4.2 K
Q = Heat to be removed to produce liquid helium
= Latent heat of vaporisation of helium
= 83.3 kJ/kmol
83.3 (305 – 4.2)
W = 4.2
= 5965.86 kJ/kmol
≈ 5966 kJ/kmol ... Ans.
Example 4.16 :
A heat exchanger is used to heat 100 kg/min of water [Cp = 4.187 kJ/(kg·K)] from 298 K
to 343 K. To meet this requirement, saturated steam at 373 K [H = 2676 kJ/kg and
S = 7.3554 kJ/(kg·K)] enters the heat exchanger and leaves as saturated liquid at 373 K
[H = 419.06 kJ/kg and S = 1.3069 kJ/(kg·K)]. Calculate the entropy change of water, steam and
the universe in one minute.
Chemical Engineering Thermodynamics - I 4.64 Second Law of Thermodynamics

Solution :
Basis : One minute of operation.
.
Water : m = 100 kg/min
∴ m = 100 × 1 = 100 kg
T1 = 298 K, T2 = 343 K and Cp = 4.187 kJ/(kg·K)
The entropy change of the water is

∆S1 = ∆Swater = mCp ln T 


T2
 1
∆S1 = 100 × 4.187 × ln 298
343
 
= 58.88 kJ/K ... Ans.
Heat to be added to water = mCp (T2 – T1)
= 100 × 4.187 (343 – 298)
= 18841.5 kJ
This heat is supplied by the steam used as a heating medium.
H of saturated steam = 2676 kJ/kg ... at 373 K
H of saturated liquid water = 419.06 kJ/kg ... at 373 K
So, to meet the heat requirement, the latent heat is used. Let m' be the kg of saturated steam
entering the exchanger.
∴ 18841.5 = m' (H of saturated steam – H of saturated liquid water)
= m' (2676 – 419.06)
∴ m' = 8.348 kg ≈ 8.35 kg
Mass of saturated steam = Mass of saturated liquid = 8.35 kg
Entropy of saturated steam = 7.3554 kJ/(kg·K)
Entropy of saturated liquid = 1.3069 kJ/(kg·K)
Initial entropy of steam, S1 = 8.35 × 7.3554 = 61.4176 kJ/K
Final entropy of steam, S2 = 8.35 × 1.3069 = 10.9126 kJ/K
The entropy change of steam is the difference between the entropy of the saturated liquid and
entropy of saturated steam.
The entropy change of the steam is
∆S2 = (∆S)steam = S2 – S1
= 10.9126 – 61.4176 = – 50.505 kJ/K ... Ans.
∆Suniverse = ∆Ssystem + ∆Ssurroundings
= ∆Swater + ∆Ssteam + ∆Ssurroundings
= 58.88 + (– 50.505) + 0
= 8.375 kJ/K … Ans.
Chemical Engineering Thermodynamics - I 4.65 Second Law of Thermodynamics

Entropy generated = 8.375 kJ/K. It is greater than zero and hence the process is
irreversible.
In this case, ∆Ssurroundings = 0 ... since there is no heat exchange between the system and
the surroundings ... true for an insulated heat exchanger.
The universe is an isolated system (for which dQ = 0).
∴ ∆Suniverse = ∆Swater + ∆Ssteam
= 58.88 + (– 50.505) = 8.375 kJ/K … Ans.
Example 4.17 :
An inventor claims to have developed a heat pump, having a coefficient of performance of 6,
which maintains a cold space at 250 K while operating in a surrounding temperature of 310 K.
Would you agree with his claim ?
Solution :
Our aim is to maintain the cold space at 250 K when the temperature of surroundings is
310 K. Heat is to be rejected from the cold space (low temperature) to the surroundings
(high temperature).
So, this heat pump is operating in reverse, i.e., as a refrigerating machine (heat pump
operating in summer).
T2 = Temperature of cold space = 250 K ... low temperature
T1 = Temperature of surroundings = 310 K ... high temperature
The coefficient of performance of an ideal refrigerating machine is given by
T2
COP = T – T
1 2

250
= 310 – 250 = 4.167

No machine operating between 250 K and 310 K can have a COP greater than 4.167. Hence,
the inventor's claim of a COP of 6 is unacceptable. … Ans.
Example 4.18 :
Suppose a hot body is available at 473 K while the ambient atmosphere is at 298 K.
Calculate the maximum efficiency of a heat engine that operates between these two bodies.
Calculate the power delivered by the engine, if it absorbs heat at the rate of 10 kJ/s from the hot
body.
Solution :
Source/hot body temperature = T1 = 473 K
Sink/atmospheric temperature = T2 = 298 K
The maximum efficiency of an engine operating between these temperatures is the Carnot
efficiency.
Chemical Engineering Thermodynamics - I 4.66 Second Law of Thermodynamics

The efficiency of a Carnot engine is given by


T1 – T2
η = T1
473 – 298
= 473 = 0.37 or 37%
The maximum efficiency of the engine is 0.37. ... Ans.
Q1 = Heat absorbed by the heat engine from the hot body = 10 kJ/s
Power delivered by the engine is given by
W = ηQ1
= 0.37 × 10 = 3.7 kJ/s = 3.7 kW ... Ans.
Example 4.19 :
An inventor claims that his heat engine has the following specifications :
(i) Power developed = 70 kW
(ii) Fuel burnt = 4 kg/h
(iii) Heating value of fuel = 75000 kJ/kg
(iv) Temperature limits = 1000 K and 300 K
State whether this claim is valid.
Solution :
Temperature limits : T1 = 1000 K (source) and T2 = 300 K (sink). The maximum possible
efficiency is given by the Carnot engine.
The efficiency of a Carnot engine is given by
T1 – T2
η = T1
1000 – 300
= 1000 = 0.70 or 70% ... maximum efficiency
Power developed by the engine = 70 kW = 70 kJ/s
Heat supplied/absorbed = Heat developed by burning fuel
= Fuel burnt × Calorific value of fuel
4 kg/h
= 3600 s/h × 75000 kJ/kg
Q1 = 83.33 kJ/s
The actual efficiency/the efficiency claimed is
W
ηactual = Q
1
70
= 83.33 = 0.84 or 84%
∴ ηactual or ηclaimed > ηcarnot engine
The inventor's claim is not valid (not feasible) since the efficiency of the engine claimed
(0.84) is greater than that of a Carnot engine (0.70). ... Ans.
Chemical Engineering Thermodynamics - I 4.67 Second Law of Thermodynamics

Example 4.20 :
The efficiency of a Carnot engine can be increased either by increasing the source
temperature (T1) by ∆T keeping the sink temperature (T2) constant or by decreasing the sink
temperature (T2) by ∆T keeping the source temperature (T1) constant. Determine which of the
above alternatives is better.
Solution :
T1 = Temperature of the source
T2 = Temperature of the sink
The efficiency of a Carnot engine operating between the above temperature limits is given by
T1 – T2
η = T1 , T1 > T2
(i) Let that the source temperature be increased by ∆T and the sink temperature be kept
constant at T2.
Source temperature = T1 + ∆T
Sink temperature = T2
(T1 + ∆T) – T2 T1 – T2 + ∆T
The efficiency is ηA = = ... (1)
T1 + ∆T T1 + ∆T
(ii) Let that the sink temperature be decreased by ∆T and the source temperature be kept
constant at T1.
Source temperature = T1
Sink temperature = T2 – ∆T
The efficiency is
T1 – (T2 – ∆T) T1 – T2 + ∆T
ηB = T1 = T1 ... (2)

Comparing Equations (1) and (2), we see that the numerators of both equations are the same
and the denominator of Equation (2) is smaller than that of Equation (1) for any +ve value of ∆T.
∴ ηB > ηA ... for any T1, T2 and ∆T values
∴ The second alternative of decreasing the sink temperature keeping the source
temperature constant is better to increase η. ... Ans.
Example 4.21 :
A Carnot refrigerator consumes 200 W power in summer when the ambient atmosphere is at
313 K. The heat losses from the walls of the cold space are estimated at 15 W per degree Celsius
temperature difference between the ambient and the cold space. Find the temperature of the cold
space.
Solution :
Let T2 = be the temperature of the cold space.
T1 = Temperature of the ambient atmosphere
T1 = 313 K
W = Power consumed by the refrigerator
= 200 W
Chemical Engineering Thermodynamics - I 4.68 Second Law of Thermodynamics

Heat loss from the walls of cold space = 15 W per degree Celsius temperature difference
Q2 = Heat absorbed from the cold space by the refrigerator
= Heat loss from the cold space
= 15 × (T1 – T2)
= 15 × (313 – T2) … in W
The coefficient of performance of a Carnot refrigerator is given by
T2 Q2
COP = T – T = W
1 2

T2 15 (313 – T2)
∴ 313 – T2 = 200
∴ 200 T2 = (313 – T2) (4695 – 15 T2)
2
∴ 15 T2 – 9590 T2 + 1469535 = 0
[(– 9590)2 – 4 × 15 × 1469535]1/2
∴ T2 = – (– 9590) ±
2 × 15
(9590 ± 1948.3)
= 30 = 384.61 or 254.7 K
T2 = 384.61 K is not the feasible root as it is higher than 313 K.
∴ T2 = 254.7 K ... is the feasible root as T of the cold space should be less than T of
refrigerator. Temperature of the cold space = 254.7 K ... Ans.
Example 4.22 :
The COP of a refrigerator can be increased either by decreasing the hot body temperature
(T1) by ∆T keeping the cold body temperature (T2) constant or increasing the cold body
temperature (T2) by ∆T keeping the hot body temperature constant. Determine which of the
above alternatives is more effective.
Solution :
T1 = Temperature of the hot body
T2 = Temperature of the cold body
The coefficient of performance of a refrigerator in terms of the temperatures of the bodies
between which it operates is given by
T2
COP = T – T
1 2

Case 1 : Let the hot body temperature be decreased by ∆T.


∴ Temperature of the hot body = T1 – ∆T
Temperature of the cold body = T2 ... held constant
For this situation, the COP is
T2 T2
(COP)1 = = ... (1)
(T1 – ∆T) – T2 T1 – T2 – ∆T
Chemical Engineering Thermodynamics - I 4.69 Second Law of Thermodynamics

Case 2 : Let the cold body temperature be increased by ∆T, keeping the hot body
temperature constant.
Temperature of the cold body = T2 + ∆T
Temperature of the hot body = T1 … held constant
For this situation, the COP is
T2 + ∆T T2 + ∆T
(COP)2 = = ... (2)
T1 – (T2 + ∆T) T1 – T2 – ∆T
Comparing Equations (1) and (2), we see that the denominators of Equations (1) and (2) are
the same, while the numerator of Equation (2) is higher than that of Equation (1).
∴ (COP)2 will be always higher than (COP)1.
∴ The second alternative of increasing the cold body temperature by ∆T keeping the hot
body temperature constant is more effective to get a high COP. ... Ans.
Example 4.23 :
A heat engine receives 600 kJ heat from a source at 1000 K during a cycle. It converts
150 kJ of this heat to net work and rejects the remaining 450 kJ to a sink at 300 K. Judge
whether this engine violates the second law of thermodynamics or not on the basis of :
(i) the Clausius inequality,
(ii) the Carnot principle.
Solution :

Source
at T1 = 1000 K

Q1 = 600 kJ

Heat
Engine W = 150 kJ

Q2 = 450 kJ

Sink at
T2 = 300 K

Fig. E 4.23
T1 = Source temperature, T2 = Sink temperature
Q1 = 600 kJ = Heat absorbed by the engine from the source at 1000 K
Q2 = 450 kJ = Heat rejected by the engine to the sink at 300 K
W = 150 kJ = Work done by the engine
Chemical Engineering Thermodynamics - I 4.70 Second Law of Thermodynamics

According to the first law of thermodynamics,


Q1 – Q2 = W
600 – 450 = 150
(i) The Clausius inequality is
dQ

O T ≤ 0

dQ
If ⌠

O T = 0 ... cycle is reversible

dQ
If ⌠

O T < 0 ... cycle is irreversible

dQ
and if ⌠

O T > 0 ... cycle is impossible

dQ Q1 Q2
∴ ⌠

O T = T –T
1 2

600 450
= 1000 – 300 = – 0.90 ... less than zero (< 0)

dQ
As ⌠⌡
O T is less than zero, the cycle is irreversible, so the engine does not violate the second
law of thermodynamics. ... Ans.
As the Clausius inequality is satisfied, the engine is theoretically possible.
(ii) The Carnot principle :
ηactual engine < ηCarnot engine ... for fixed T1 and T2
The efficiency of a Carnot engine is given by
T1 – T2 1000 – 300
ηCarnot = T1 = 1000 = 0.70

The efficiency of the actual engine is


W Q1 – Q2 600 – 450
ηactual engine = Q = Q = 600 = 0.25
1 1

∴ ηactual (0.25) < ηCarnot (0.70) ... possible /feasible engine


As the efficiency of the given heat engine is less than that of a Carnot engine, this heat
engine does not violate the second law of thermodynamics. ... Ans.
Example 4.24 :
A heat pump is used to maintain a house at 298 K when the ambient atmosphere is at 283 K.
The house is losing heat to the ambient atmosphere through the walls and windows at a rate of
60000 kJ/h while the heat generated within the house from the people, lights and appliances
amounts to 4000 kJ/h. Determine the power input to the heat pump if the COP of the actual heat
pump is 30% of the COP of a Carnot heat pump.
Chemical Engineering Thermodynamics - I 4.71 Second Law of Thermodynamics

Solution :

House at T1 = 298 K

Q1

Heat
W Pump

Q2

Ambient atmosphere at
T2 = 283 K

Fig. E 4.24
T1 = 298 K and T2 = 283 K
The COP of an ideal heat pump is given by
T1
(COP)Carnot HP = T – T
1 2

298
= 298 – 283 = 19.867

Given : COP of the actual heat pump = 30% or 0.30 of COP of the Carnot heat pump.

The COP of the actual heat pump is

(COP)HP = 0.30 × 19.867 = 5.96

Heat loss from the house = 60000 kJ/h = 16.67 kJ/s

Heat generated in the house = 4000 kJ/h = 1.11 kJ/s

Net heat loss from the house = 16.67 – 1.11 = 15.56 kJ/s

For the house to maintain at 298 K, we have to supply 15.56 kJ/s heat to the house.
Q1 = Heat rejected to the house = 15.56 kJ/s

The COP of the heat pump is


Q1
COP = W

Q1 15.56
∴ W = COP = 5.96 = 2.61 kJ/s = 2.61 kW

Power input to the heat pump = 2.61 kW ... Ans.


Chemical Engineering Thermodynamics - I 4.72 Second Law of Thermodynamics

Example 4.25 :
A refrigerator is to be used to maintain a cold storage at 250 K (– 23°C). The compressor is
rated at 250 W and the cooling duty is estimated at 50000 kJ/day. Determine the fraction of the
time the compressor runs in a whole day if the refrigerator is used in
(i) a tropical country where the ambient temperature is 310 K and
(ii) a cold country where the ambient temperature is 290 K.
Solution :
(i) Refrigerator used in a tropical country where the ambient temperature is 310 K.
T2 = 250 K and T1 = 310 K
The COP of a refrigerator is given by
T2 Q2
COP = T – T = W
1 2
250
COP = 310 – 250 = 4.167 ... maximum COP
Cooling duty = 50000 kJ/day ... refrigerating effect
Q2 = Heat absorbed from the cold storage
= Cooling duty
= 50000 kJ/day
Let us obtain the compressor work required for the cooling duty of 50000 kJ/day.
Q2 50000
W = COP = 4.167 = 11999 kJ/day
The compressor work required for the given cooling duty/refrigerating effect is
11999 kJ/day.
The compressor is rated at 250 W = 250 J/s.
∴ If the compressor is run for 24 hours, then the work that can be performed by the
compressor is
W' = 250 J/s × (24 × 3600 s)
= 21600000 J = 21600 kJ
[This is a capacity of the compressor to perform work in a day.]
 Fraction of the time  Compressor work for the given cooling duty
the compressor works =
 in a whole day   Work that can be performed by 
the compressor if it runs for 24 h
W 11999
= W' = 21600 = 0.555 ≈ 0.56 ... Ans. (i)
[i.e., the compressor runs for 0.56 × 24 = 13.44 h in a whole day.]
(ii) Refrigerator used in a cold country where the ambient temperature is 290 K :
T2 = 250 K
T1 = 290 K
Chemical Engineering Thermodynamics - I 4.73 Second Law of Thermodynamics

The COP of the refrigerator is


T2
COP = T – T
1 2
250
= 290 – 250 = 6.25
Heat to be absorbed from the cold storage = Cooling duty
Q2 = Refrigerating effect = 50000 kJ/day
i.e., in 24 hours, the heat absorbed from the cold storage by the refrigerator = 50000 kJ
Let us obtain the compressor work required for the above cooling duty.
Q2 50000
W = COP = 6.25 = 8000 kJ ... per day
Our compressor is rated as 250 W = 200 J/s
i.e., our compressor performs a work of 200 J per second.
If this compressor is run for 24 hours, then the work that can be performed by it is
W' = 200 J/s (24 × 3600 s)
= 21600000 J = 21600 kJ
 Fraction of the time  Compressor work for the given cooling duty
the compressor works =
 in a whole day  Work that can be performed by the
 compressor if it runs for 24 h 
8000
= 21600 = 0.37 ... Ans. (ii)

[i.e., the compressor runs for 0.37 × 24 = 8.88 h in a whole day.]


Example 4.26 :
A Carnot engine operating between a source at 1073 K and the ambient atmosphere at 298 K
is used to operate a Carnot refrigerator between 253 K and 298 K. If the engine absorbs 10 kJ/s
from the source at 1073 K, determine (i) the rate at which the refrigerator can absorb energy as
heat from the low temperature body and (ii) the rate of heat rejection to the ambient atmosphere
at 298 K by both the devices.
Solution :

Fig. E 4.26 : Schematic diagram of a refrigerator supported by a heat engine


Chemical Engineering Thermodynamics - I 4.74 Second Law of Thermodynamics

(1) Carnot engine :


T1 = 1073 K ... Source temperature
T2 = 298 K ... Sink temperature ... Ambient atmosphere
The efficiency of a Carnot engine is given by
T1 – T2 1073 – 298
η = T1 = 1073 = 0.72

Energy absorbed by the engine = Q1 = 10 kJ/s


Power developed by the heat engine is
W = ηQ1 = 0.72 × 10 = 7.2 kJ/s
Power delivered to the refrigerator = W = 7.2 J/s
(2) Carnot refrigerator :
T2 = 253 K ... Temperature of the cold body
T1 = 298 K ... Temperature of the ambient atmosphere
The coefficient of performance of a Carnot refrigerator is given by
T2 253
COP = T – T = 298 – 253 = 5.62
1 2

The rate of heat absorption by the refrigerator is


Q2R = COP × W
= 5.62 × 7.2 = 40.464 kJ/s ≈ 40.46 kJ/s ... Ans. (i)
[It is the capacity of the given refrigerator to extract heat from the cold space.]
The subscript R in Q2R, Q1R stands for refrigerator.
Heat rejected by the engine to the ambient atmosphere is
Q2 = Q1 – W
= 10 – 7.2 = 2.8 kJ/s
Heat rejected by the refrigerator to the ambient atmosphere is
Q1R = Q2R + W
= 40.46 + 7.2 = 47.66 kJ/s
The rate of heat rejection to the ambient atmosphere by both the devices is
Q2 + Q1R = 2.8 + 47.66 = 50.46 kJ/s ... Ans. (ii)
[We have, Q2 + Q1R = Q1 + Q2R = 10 + 40.46 = 50.46 kJ/s]
Example 4.27 :
It is desired to produce 5000 kg/h of ice at 273 K (0°C) from water at 273 K, while the
ambient temperature is 313 K. It is planned to supply power from a heat engine to run the
refrigerator. The heat engine operates between a source at 373 K and the ambient atmosphere.
Chemical Engineering Thermodynamics - I 4.75 Second Law of Thermodynamics

Calculate (i) the minimum power required to run the refrigerator, (ii) the maximum
efficiency of the heat engine, and (iii) the ratio of heat rejected by both the devices to the ambient
atmosphere to the energy absorbed by the refrigerator from the water at 273 K.
Data : The latent heat of fusion of water at 273 K is 6.002 kJ/mol.
Solution :

Fig. E 4.27 : A refrigerator supported by a heat engine


T2 = Temperature of cold body = 273 K
T1 = Temperature of hot body (ambient atmosphere) = 313 K
The work (power) required will be minimum for an ideal Carnot machine acting as a
refrigerator. The coefficient of performance of a Carnot refrigerator is given by
T2 273
COP = T – T = 313 – 273 = 6.825
1 2

·
Rate of ice production from water at 273 K = m = 5000 kg/h
Let us convert the latent heat from kJ/mol to kJ/kg.
Latent heat of fusion of water at 273 K = 6.002 kJ/mol = 6.002 × 103 kJ/kmol
6.002 × 103 kJ/kmol
λ = 18 kg/kmol = 333.44 kJ/kg

∴ For freezing 1 kg of water, the heat to be removed from it is 333.44 kJ


[Molecular weight of H2O = 18 kg/kmol]
Rate of heat absorption by the refrigerator from water at 273 K = Q2R
The subscript R in Q1R and Q2R stands for refrigerator.
Rate of heat removal from water by the refrigerator is
·
Q2R = m λ = 333.44 kJ/kg × 5000 kg/h
1667200 kJ/h
= 1667200 kJ/h = 3600 s/h = 463.11 kJ/s = 463.11 kW
Chemical Engineering Thermodynamics - I 4.76 Second Law of Thermodynamics

The minimum power required to run the refrigerator is


Q2R 463.11
W = COP = 6.825 = 67.85 kW ... Ans. (i)

This power is supplied by the heat engine.


Power developed by the heat engine and delivered to the refrigerator is
W = 67.85 kW
The maximum efficiency of an engine operating between 373 K and 313 K is the Carnot
efficiency.
The efficiency of a Carnot engine is given by
T1 – T2 373 – 313
η = T1 = 373 = 0.1608 ≈ 0.161 or 16.1%

The maximum possible efficiency of the heat engine is 0.161 ... Ans. (ii)
Heat energy absorbed by the heat engine from the source at 373 K = Q1
W
We have, η = Q
1

W 67.85
∴ Q1 = = 0.161 = 421.43 kW
η
Heat rejected by the heat engine to the ambient atmosphere is
Heat rejected = Heat absorbed – Power developed and delivered
Q2 = Q1 – W
= 421.43 – 67.85 = 353.58 kW
[since Q1 = Q2 + W ... from the first law of thermodynamics.]
Heat rejected by the refrigerator to the ambient atmosphere is
Heat rejected = Heat absorbed from cold body + Power consumed
Q1R = Q2R + W = 463.11 + 67.85 = 530.96 kJ
∴ Heat rejected to the ambient atmosphere by both the devices
= Q2 + Q1R
= 353.58 + 530.96 = 884.54 kW
[Q1 + Q2R = 421.43 + 463.11 = 884.54 kW = Q2 + Q1R]
The ratio of the heat rejected to the ambient atmosphere to the heat absorbed from the water
at 273 K by the refrigerator is
Q2 + Q1R 884.54
Q2R = 463.11 = 1.9099 ≈ 1.91 ... Ans. (iii)
Chemical Engineering Thermodynamics - I 4.77 Second Law of Thermodynamics

Example 4.28 :
It is proposed to maintain a drawing hall at 298 K (25°C) in summer as well as in winter.
In summer, the ambient temperature rises to as high as 313 K (40°C) and in winter it drops to
as low as 276 K (3°C). The heat loss through the walls and roofing is estimated to be 20 kJ/s per
degree temperature difference between the ambient and the drawing hall. Calculate the minimum
power required to maintain the hall at 298 K in summer as well as in winter if the same device is
used as a refrigerator in summer and as a heat pump in winter.
Solution :
Ambient atmosphere Ambient atmosphere
at T1 = 313 K at T2 = 276 K

Q1 Q2

W W Heat
Refrigerator pump

Q2 Q1

Drawing Hall Drawing Hall


at T2 = 298 K at T1 = 298 K

(A) Summer (B) Winter


Fig. E 4.28 : (A) Device operating as a refrigerator in summer
(B) Device operating as a heat pump in winter
The power required will be minimum for the Carnot machine (reversible machine)
operating/acting as a refrigerator or heat pump.

(1) Refrigerator (for summer) :


T2 = 298 K ... Temperature of the drawing hall (cold body)

T1 = 313 K ... Temperature of the ambient atmosphere

In summer, heat has to be removed from the drawing hall and delivered to the ambient
atmosphere to maintain the hall at 298 K. Here, heat will leak/flow into the hall from the
atmosphere which is at higher temperature (313 K).
Q2 = Heat to be removed/absorbed by the refrigerator

= Heat flow from outside through the walls and roofing


= Heat loss per degree temperature difference) × (T1 – T2)

= 20 kJ(s.K) × (313 – 298) K = 300 kJ/s


Chemical Engineering Thermodynamics - I 4.78 Second Law of Thermodynamics

The coefficient of performance of a Carnot refrigerator (an ideal refrigerator) is given by


T2
(COP)R = T – T
1 2

298
= 313 – 298 = 19.87 ... (COP)max

The minimum power required by the refrigerator is


Q2 Q2
(W)R = (COP) = (COP)
R max

300
= 19.87 = 15.098 ≈ 15.1 kJ/s = 15.1 kW ... Ans.

(2) Heat pump (for winter) :


T1 = 298 K ... Temperature of the drawing hall (hot body)
T2 = 276 K ... Temperature of the ambient atmosphere (cold body)

In winter, the ambient atmosphere is at 276 K and the hall is at 298 K, so heat will leak/flow
from the hall to the atmosphere and the heat pump will supply the same amount of heat to
maintain the hall at 298 K.
QH = Heat to be supplied by the heat pump

= Heat that flows from the walls and roofing of the hall to the ambient
atmosphere (Heat loss per degree temperature difference)
= 20 kJ/(s·K temperature difference) × (T1 – T2) K
= 20 (298 – 276) = 440 kJ/s
The coefficient of performance of a Carnot heat pump is given by
T1
(COP)HP = T – T
1 2

298
= 298 – 276 = 13.545 ... (COP)max

The minimum power required by the heat pump is


Q1 440
(W)HP = (COP) = 13.545
HP

= 32.48 kJ/s = 32.48 kW ... Ans.

[Since the temperature of the ambient atmosphere is different in the summer (313 K) than
that in the winter (276 K), COP (heat pump) = 1 + COP (refrigerator) is not valid for this case.]
Chemical Engineering Thermodynamics - I 4.79 Second Law of Thermodynamics

Example 4.29 :
A heat pump is used to heat a house in winter and then reversed to cool the house in summer.
The house is to be maintained at 298 K (25°C). Heat transfer through the walls and roof is
estimated to be 1890 kJ per hour per degree temperature difference between the house and the
ambient atmosphere.
(i) If the ambient temperature is 275 K (2°C) in winter, determine the minimum power
required to drive the heat pump.
(ii) For the same power input, in summer, determine the maximum ambient temperature for
the house temperature at 298 K.
Solution :
Ambient atmosphere
at T2 = 275 K

Q2

W Heat
pump

Q1

House
Q
at T1 = 298 K

Fig. E 4.29 (A) : Device working as a heat pump in winter


(i) Heat pump (in winter) : In winter, the house is at 298 K (higher temperature) and the
ambient atmosphere is at 275 K (lower temperature). Hence, the heat will flow out of the house
and the same amount of heat should be supplied to the house to maintain it at 298 K.
Heat transfer through the walls and roof = 1890 kJ per hour per degree temperature
difference.
Q1 = Heat rejected to the house (high temperature body)
= Heat transfer × Temperature difference
= 1890 kJ/(s·K temperature difference) × (T1 – T2) K
= 1890 × (298 – 275) = 43470 kJ/h = 12.075 kJ/s = 12.075 kW
The coefficient of performance of a Carnot heat pump is given by
T1
(COP)HP = T – T ... maximum COP
1 2

298
= 298 – 275 = 12.956
Chemical Engineering Thermodynamics - I 4.80 Second Law of Thermodynamics

The power required by a heat pump will be minimum when the COP of the heat pump is
maximum.
The minimum power required to drive the heat pump is
Q1 12.075
W = (COP) = 12.956 = 0.932 kW ... Ans. (i)
HP

(ii) Refrigerator (in summer) :


Ambient
atmosphere at T1

Q1

W
Refrigerator

Q2

House at
T2 = 298 K

Fig. E 4.29 (B) : Device working as a refrigerator in summer


In summer, heat will have to be removed from the house and delivered to the ambient
atmosphere to maintain the house at 298 K which is less than TH (or T1) - ambient temperature.
Heat will flow into the house from the ambient atmosphere which is at higher temperature
and this much quantity of heat has to be removed.
Q2 = Heat absorbed from the house by the refrigerator
= 1890 kJ/(s·K temperature difference) × Temperature difference
= 1890 × (T1 – T2)
= 1890 × (T1 – T2) = 1890 (T1 – 298) kJ/h
= 0.525 (T1 – 298) kJ/s = 0.525 (T1 – 298) kW
where T1 is the temperature of the ambient atmosphere in K.
The power required by a refrigerator is minimum when it has a maximum COP.
Given : Power input of the heat pump = Power input of the refrigerator
∴ Power required by the refrigerator = W = 0.932 kW
The COP of an ideal refrigerator is given by
T2
(COP)R = T – T
1 2

298
(COP)R = T – 298 ... maximum COP … (1)
1

Q2
(COP)R = W … (2)
Chemical Engineering Thermodynamics - I 4.81 Second Law of Thermodynamics

Equating Equations (1) and (2), we get


298 Q2 0.525 (T1 – 298)
T1 – 298 = W = 0.932
298 × 0.932 = (0.525 T1 – 0.525 × 298) (T1 – 298)
2
∴ 0.525 T1 – 312.9 T1 + 46344.36 = 0
– (– 312.9) ± [(– 312.9)2 – 4 × (0.525) × (46344.36)]1/2
∴ T1 =
2 × 0.525
(312.9 ± 24.15)
= 1.05 = 321 or 275 K
In summer, temperature of house (298 K) to be maintained is less than that of the ambient
atmosphere. Hence, temperature of 275 K is not the correct root as it is less than 298 K.
The maximum ambient temperature for W = 0.932 kW and house at 298 K is
321 K (i.e., 48°C). ... Ans.
Example 4.30 :
An inventor claims to have developed a cyclically operating device that absorbs 500 kJ and
300 kJ heat from reservoirs at 800 K and 400 K respectively and rejects 100 kJ and 50 kJ heat to
reservoirs at 600 K and 300 K respectively, while it delivers 650 kJ work. Would you agree with
his claim ?
Solution :
Thermal Thermal
Reservoir Reservoir
T1 = 800 K T2 = 400 K

Q1 = 500 kJ Q2 = 300 kJ

Heat
Engine W = 650 kJ

Q3 = 100 kJ Q4 = 50 kJ
Thermal Thermal
Reservoir Reservoir
T3 = 600 K T4 = 300 K

Fig. E 4.30 : Heat engine operating with four thermal reservoirs


Consider a heat engine as shown in Fig. E 4.30.
The first law of thermodynamics gives
Q1 + Q2 – Q3 – Q4 = W
500 + 300 – 100 – 50 = 650
650 = 650 ... satisfies the first law
Chemical Engineering Thermodynamics - I 4.82 Second Law of Thermodynamics

The second law of thermodynamics in the form of Clausius inequality is


dQ

O T ≤ 0
⌡ … inequality holds good for irreversible cycles and
equality holds good for reversible cycles
If ⌠
O T > 0‚ then the cycle is impossible (the cyclic process is not feasible).
dQ
 ⌡ 
Now, we have to check whether the Clausius inequality can be satisfied or not. If it satisfies
the Clausius inequality, then the claim is accepted.
dQ Q1 Q2 Q3 Q4

⌡O T = T +T –T –T
1 2 3 4

500 300 100 50


= 800 + 400 – 600 – 300

= 0.625 + 0.75 – 0.167 – 0.167


= 1.041 > 0
dQ
In our case, ⌠

O T > 0 (1.041), which is in violation of the thermodynamics (i.e., the claim
violates the second law of thermodynamics) and hence the claim is invalid. I am not agree with
the inventor's claim as it violates the second law of thermodynamics. ... Ans.
No. ... violates the second law of thermodynamics. ... Ans.
Example 4.31 :
A domestic refrigerator which maintains a low temperature of 258 K in the deep freezer and
capable of absorbing 200 J/s energy as heat from the low temperature body is imported from
abroad, where the ambient temperature is 298 K. If the same refrigerator is put into service in a
tropical country, where the ambient temperature is 313 K and the thermostat is set at 258 K and
the power input to the refrigerator is the same as that in the cold country, estimate the rate at
which it can absorb energy from the low temperature body.

Solution :

Case I : Refrigerator used in a cold country :


Temperature of the cold body = T2 = 258 K

Temperature of the ambient atmosphere = T1 = 298 K

The coefficient of performance of a refrigerator is given by


T2
[COP]I = T – T
1 2

258
= 298 – 258 = 6.45
Chemical Engineering Thermodynamics - I 4.83 Second Law of Thermodynamics

The power input (W)I is given by


[Q2]I
[W]I = [COP]
I

where [Q2]I = Heat absorbed from the cold body by the refrigerator
= 200 J/s
200
∴ [W]I = 6.45 = 31 J/s
Case II : Refrigerator used in a tropical country
T2 = 258 K and T1 = 313 K
Given : [Power input, i.e., W]II = [Power input, i.e., W]I
∴ [W]II = 31 J/s
T2
[COP]II = T – T
1 2
258
= 313 – 258 = 4.691
The rate of heat absorption by the refrigerator from the cold body in the tropical country is
given by
[Q2]II = [COP]II × [W]II
= 4.691 × 31 = 145.42 J/s ... Ans.
Example 4.32 :
It is desired to maintain a large cold storage at 283 K when the atmospheric temperature is
313 K. The energy losses as heat from the walls of the cold storage are 10 kJ/s per degree kelvin
difference between the atmosphere and the cold space. What is the minimum power required to
maintain the cold storage at 283 K ?
Solution :
The power required will be minimum for an ideal Carnot engine acting as a refrigerating
machine, i.e., when the COP of a refrigerating machine is maximum.
Temperature of the cold storage = T2 = 283 K
Temperature of the surroundings = T1 = 313 K
The energy losses as heat from the walls of the cold storage are 10 kJ/s per degree kelvin
temperature difference between the atmosphere and the cold space.
∴ Q2 = 10 kJ/(s·K) × Temperature difference (T1 – T2), K
= 10 × (313 – 283)
= 300 kJ/s = 300 kW
The coefficient of performance of an ideal refrigerating machine (or an ideal Carnot engine
as a refrigerating machine) is given by
T2 283
COP = T – T = 313 – 283 = 9.433
1 2
Chemical Engineering Thermodynamics - I 4.84 Second Law of Thermodynamics

Q2
We have, COP = W

Q2 300
∴ W = COP = 9.433 = 31.80 kW

The minimum power required to maintain the cold storage at 283 K is 31.80 kW ... Ans.
Example 4.33 :
The work output of an ideal Carnot engine operating between two thermal reservoirs at
1000 K and 300 K respectively, is utilized to run the compressor of a refrigerator working on
R-12. The heat rejected by the engine to the sink is 30 kW. The refrigerator operates between
240 K and 300 K.
Calculate : (i) the capacity of the refrigerator, (ii) the COP of the refrigerator and (iii) the
circulation rate of the refrigerant.
Data : The enthalpy of saturated Freon-12 (R-12) liquid at 300 K = 61.90 kJ/kg and the
enthalpy of saturated R-12 vapour at 240 K = 172.8 kJ/kg.
Solution :

Source at Q1 Carnot Q2 Sink at


T1 = 1000 K Engine T2 = 300 K

Cold space at Ambient atmosphere


Refrigerator
T2 = 240 K Q2R Q1R at T1 = 300 K

Fig. E 4.33 : Refrigerator supported by a Carnot engine


The efficiency of an ideal Carnot engine is given by
W T1 – T2
η = Q = T2
1

where T1 = 1000 K and T2 = 300 K


W 1000 – 300
Q1 = 1000 = 0.70

∴ W = 0.70 Q1
Given : Heat rejected to the sink at 300 K by the engine = Q2 = 30 kW
We have, Q1 – Q2 = W
Q1 – 30 = 0.70 Q1
∴ Q1 = 100 kW
W = 0.70 Q1 = 0.70 × (100) = 70 kW
Chemical Engineering Thermodynamics - I 4.85 Second Law of Thermodynamics

Power developed by the engine and delivered to the refrigerator = W = 70 kW


The COP of a refrigerator is given by
Q2R T2
[COP]R = W = T – T
1 2

where T2 = 240 K and T1 = 300 K


The subscript R in Q2R stands for refrigerator.
240
∴ COP = 300 – 240 = 4

The COP of the refrigerator is 4. ... Ans. (ii)


Q2R
W = COP
∴ Q2R = COP × W = 4 × 70 = 280 kW
Q2R = Heat absorbed from the cold space by the refrigerant
= Capacity of the refrigerator
= 280 kW ... Ans. (i)
Enthalpy of saturated vapour R-12 at 240 K = H1 = 172.8 kJ/kg
Enthalpy of saturated liquid R-12 at 300 K = H2 = 61.90 kJ/kg
·
Let m be the circulation rate of refrigerant in kg/s.
.
Q2R = m (H2 – H1) ... cold space energy balance
[Heat absorbed during vaporisation is equal to the change in enthalpy of the refrigerant]
where Q2R = 280 kW = 280 kJ/s
· Q2R 280
m = H – H = 172.8 – 61.90 = 2.52479 ≈ 2.525 kg/s
2 1

The circulation rate of the refrigerant is 2.525 kg/s. ... Ans. (iii)
Example 4.34 :
A vapour compression refrigerating unit with Freon-12 as refrigerant works between
pressure limits of 182.5 kPa and 960.6 kPa. The heat absorbed in the evaporator is 3200 kJ/h
and that rejected in the condenser is 72 kJ/min. The refrigerant vapour leaves the evaporator in
the saturated state. Calculate :
(i) the energy input to the compressor,
(ii) the refrigerant circulation rate in kg/min and
(iii) the COP of the system.
Data : H of saturated vapour at 182.5 kPa = 181.2 kJ/kg
and H of saturated liquid at 960.6 kPa = 76.2 kJ/kg
Chemical Engineering Thermodynamics - I 4.86 Second Law of Thermodynamics

Solution :
Heat absorbed in the evaporator = Q2 = 3200 kJ/h = 53.33 kJ/min
Heat rejected in the condenser = Q1 = 72 kJ/min
W = Power required by (energy input to) the compressor
= Q1 – Q2
= 72 – 53.33 = 18.67 kJ/min
18.67
= 60 = 0.311 kJ/s = 0.311 kW ... Ans. (i)
Q2 = 53.33 kJ/min = 0.889 kJ/s
The COP of the refrigerating unit is
Q2 0.889
COP = W = 0.311 = 2.858 ≈ 2.86 ... Ans. (iii)
.
Let m be the circulation rate of the refrigerant in kg/min.
Heat absorbed in the evaporator = Q2 = 53.33 kJ/min
i.e., it is the heat removed from the cold space and absorbed by the refrigerant.
Heat absorbed during vaporisation = Change in the enthalpy of refrigerant
.
Q2 = m (H2 – H1)
where H2 = H of saturated vapour at 182.5 kPa = 181.2 kJ/kg
H1 = H of saturated liquid at 960.6 kPa = 76.2 kJ/kg
. Q2 53.33
m = (H – H ) = (181.2 – 76.2)
2 1

= 0.5079 kg/min ≈ 0.508 kg/min


The refrigerant (Freon-12) circulation rate is 0.508 kg/min. ... Ans. (ii)
Example 4.35 :
A refrigerator requires 1 kW of power for a refrigeration rate of 3 kJ/s.
Calculate : (i) the COP of the refrigerator, (ii) the heat rejected and (iii) the lowest
temperature that can be maintained in the cold space of the refrigerator if the heat is rejected at
308 K.
Solution :
Heat absorbed by the refrigerator = Q2 = 3 kJ/s
Power/work required/power input = W = 1 kW = 1 kJ/s
The heat rejected is
Q1 = Q2 + W = 3 + 1 = 4 kJ/s = 4 kW ... Ans. (ii)
The COP of the refrigerator is
Q2 3
COP = W = 1 = 3 ... Ans. (i)
Temperature at which heat is rejected = T1 = 308 K
Chemical Engineering Thermodynamics - I 4.87 Second Law of Thermodynamics

Let T2 be the lowest temperature that can be maintained in the cold space (at which the heat
is absorbed).
Q1 T1
We have, Q2 = T2
4 308
3 = T2
∴ T2 = 231 K ... Ans. (iii)
T2
or COP = T – T
1 2

T2
3 = 308 – T
2

∴ T2 = 231 K … Ans.
Example 4.36 :
Air (ideal gas, γ = 1.4) enters an adiabatic compressor at 298 K and 0.1 MPa and leaves at
1 MPa and 603 K. Calculate the isentropic efficiency of the compressor.
Solution :
Inlet : P1 = 0.1 MPa and T1 = 298 K
Exit : P2 = 1 MPa and T2 = 603 K
For reversible and adiabatic compression :
T'2 P2(γ–1)/γ , γ = 1.4
= P 
T 1  1
P2(γ–1)/γ = 298  1 (1.4 – 1)/1.4
T'2 = T1 P 
 1 0.10
= 575.35 K
We have, Ws = – ∆H = – (H2 – H1) = H1 – H2
(–Ws) = ∆H = H2 – H1 = Cp (T2 – T1)
(–Ws)actual = Cp (603 – 298)
(–Ws)isentropic = (∆H)isentropic

= Cp T2' – T1
[ ] = Cp (575.35 – 298)
The isentropic efficiency of the compressor is
(–Ws)isentropic
ηc = (–W )
s actual
Cp (575.35 – 298)
= Cp (603 – 298) = 0.908 ≈ 0.91 ... Ans.
Chemical Engineering Thermodynamics - I 4.88 Second Law of Thermodynamics

Example 4.37 :
Superheated steam at 3 MPa and 573 K enters an adiabatic turbine and leaves as saturated
vapour at 313 K. Calculate the rate of entropy generation in the turbine (due to irreversibilities
in it) if the steam enters at a rate of 1 kg/s.
Data : Entropy of superheated steam at 3 MPa and 573 K = 6.5422 kJ/(kg·K)
Entropy of saturated vapour at 313 K = 8.2583 kJ/(kg·K)
Solution :
According to the second law of thermodynamics, the rate of entropy generation due to
irreversibilities for an adiabatic steady flow process is given by
. .
Sgeneration = m (Se – Si)
where Si = Entropy at the inlet conditions
= Entropy of superheated steam at 3 MPa and 473 K
= 6.5422 kJ/(kg·K)
Se = Entropy at the exit conditions
= Entropy of saturated vapour at 313 K
= 8.2583 kJ/(kg·K)
.
Sgeneration = 1 (8.2583 – 6.5422) = 1.7161 kJ/(K·s) ... Ans.
Example 4.38 :
Saturated steam at 175 kPa is compressed at a rate of 1.5 kg/s to 600 kPa in a centrifugal
compressor. The compressor efficiency is 75%. Determine (i) the power required to run the
compressor and (ii) the enthalpy and temperature of the steam leaving the compressor.
Data : Saturated steam at 175 kPa : H = 2700.60 kJ/kg and S = 7.1717 kJ/(kg·K)
Superheated steam at 600 kPa and 573 K : H = 3061.6 kJ/kg and S = 7.3724 kJ/(kg·K)
Superheated steam at 600 kPa and 523 K : H = 2957.2 kJ/kg and S = 7.1816 kJ/(kg·K)
Superheated steam at 600 kPa and 473 K : H = 2850.1 kJ/kg and S = 6.9665 kJ/(kg·K)
Solution :
For an adiabatic compression :
Ws = – ∆H ... According to the first law of thermodynamics
The minimum shaft work is the isentropic work.
(Ws)isentropic = – ∆Hs
where ∆Hs is the enthalpy change associated with the isentropic compression.
For an isentropic compression :
S2 = S1
Entropy of steam leaving = Entropy of steam entering
Chemical Engineering Thermodynamics - I 4.89 Second Law of Thermodynamics

Given : S1 = 7.1717 kJ/(kg·K) and H1 = 2700.6 kJ/kg


∴ S2 = 7.1717 kJ/(kg·K)
Temperature and enthalpy of superheated steam at 600 kPa having entropy of
7.1717 kJ/(kg·K) can be obtained by interpolation.
Superheated steam at 600 kPa and 523 K : S = 7.1816 kJ/(kg·K)
Superheated steam at 600 kPa and 473 K : S = 6.9665 kJ/(kg·K)
S2 = 7.1717 kJ/(kg·K). It is less than S at 523 K and more than S at 473 K, so the temperature
of superheated steam is between 473 K and 523 K.

∆T  523 – 473  + 6.9665 = 7.1717


7.1816 – 6.9665
 
∴ ∆T = 47.69 ≈ 47.7 K
∆T = T of superheated steam (to be obtained) – 473
T of superheated steam at 600 kPa having entropy of 7.1717 kJ/(kg·K)
= 473 + 47.7 = 520.7 K
Temperature of superheated steam at 600 kPa having entropy of 7.1717 kJ/(kg·K)
= 520.7 K
Enthalpy at 600 kPa and 473 K = 2850.1 kJ/kg
Enthalpy at 600 kPa and 523 K = 2957.2 kJ/kg
Enthalpy at 600 kPa and 520.7 K is calculated by interpolation.
Enthalpy of superheated steam at 600 kPa and 520.7 K
(2957.2 – 2850.1) (520.7 – 473)
= 2850.1 + (523 – 473)
= 2952.27 kJ/kg
∴ H1 = 2700.6 kJ/kg
and H2 = 2952.27 kJ/kg
Ws = – ∆Hs
= – (H2 – H1) = – (2952.27 – 2700.6)
= – 251.67 kJ/kg
The compression efficiency is 75%.
(Ws)isentropic
η = (Ws)actual

– 251.67
(Ws)actual = 0.75 = – 335.56 kJ/kg
Chemical Engineering Thermodynamics - I 4.90 Second Law of Thermodynamics

The minus sign indicates that the work is required to be done to run the compressor.
Work required (–Ws)actual = 335.56 kJ/kg
.
We have, m = 1.5 kg/s
Power is work done per unit time.
·
Power = (–Ws) in kJ/kg × m in kg/s
Power required to run the compressor = 335.56 kJ/kg × 1.5 kg/s
= 503.34 kJ/s
= 503.34 kW ... Ans. (i)
The actual change of enthalpy of steam is
∆Hs 251.67
∆H = = 0.75 = 335.56 kJ/kg
η
Actual enthalpy of steam leaving the compressor = H1 + ∆H
= 2700.6 + 335.56
= 3036.16 kJ/kg ... Ans. (ii)
The temperature of superheated steam at 600 kPa and having enthalpy of 3036.16 kJ/kg that
leaves the compressor can be obtained by interpolation.
Example 4.39 :
Superheated steam at 6 MPa and 773 K enters a nozzle and discharges to a constant
pressure region at 2 MPa. Determine the velocity and temperature of the exit steam if the
expansion in the nozzle is isentropic.
Data : Superheated steam at 6 MPa and 773 K : H = 3422.2 kJ/kg and S = 6.8803 kJ/(kg·K)
Superheated steam at 2 MPa and 573 K : H = 3023.5 kJ/kg and S = 6.7664 kJ/(kg·K)
Superheated steam at 2 MPa and 623 K : H = 3137 kJ/kg and S = 6.9563 kJ/(kg·K)
Solution :
For an isentropic flow (i.e., reversible and adiabatic flow) through a nozzle, the second law
of thermodynamics gives
S2 = S1
Entropy at the exit = Entropy at the inlet ... of a nozzle
S1 = Entropy of superheated steam at 6 MPa and 773 K
= 6.8803 kJ/(kg·K)
∴ S2 = 6.8803 kJ/(kg·K) ... as S2 = S1
S2 = Entropy at the exit = Entropy of superheated steam at 2 MPa and T K
T is the temperature of the exit steam for which S = 6.8803 kJ/(kg·K) and it can be obtained
by interpolation.
Chemical Engineering Thermodynamics - I 4.91 Second Law of Thermodynamics

Steam at 2 MPa and 573 K : S = 6.7664 kJ/(kg·K)


Steam at 2 MPa and 623 K : S = 6.9563 kJ/(kg·K)
S2 = 6.8803 kJ/(kg·K) is in between the above S values.

S at 573 + (T – 573) ×  623 – 573  = 6.8803 = S2


6.9563 – 6.7664
 
6.7664 + (T – 573)  623 – 573  = 6.8803
6.9563 – 6.7664

 
(T – 573) = 29.989 ≈ 30
∴ T = 573 + 30 = 603 K
T of steam at which S = 6.8803 kJ/(kg·K) for steam at 2 MPa is 603 K.
The temperature of the exit steam/steam leaving the nozzle is 603 K. ... Ans.
Discharge/exit steam is at 2 MPa and 603 K.
The enthalpy of this steam is obtained by interpolation.
Steam at 2 MPa and 573 K : H = 3023.5 kJ/kg
Steam at 2 MPa and 623 K : H = 3137 kJ/kg
T = 623 K is greater than 573 K. H is increasing with T.
H of steam at 2 MPa and 603 K = 3023.5 +  623 – 573  [603 – 573]
3137 – 3023.5
 
= 3091.6 kJ/kg
H2 = H of steam at 2 MPa and 603 K = 3091.6 kJ/kg = 3091.6 × 103 J/kg
H1 = H of steam at 6 MPa and 773 K = 3422.2 kJ/kg = 3422.2 × 103 J/kg
u2 = Velocity of the exit/discharge steam
The nozzle is adiabatic. Hence, Q = 0.
The nozzle is horizontal. Hence, z1 = z2 and ∆z = 0.
In case of a nozzle, no shaft work is delivered. Therefore, Ws = 0.
Negligible inlet velocity. Hence, u1 = 0
The first law of thermodynamics for a steady flow process for the above conditions reduces
to
2
u2
(H2 – H1) + 2 = 0 ... in J/kg
2
u2
∴ 2 = H1 – H2
∴ u2 = [2 (H1 – H2)]1/2
= [2 (3422.2 × 103 – 3091.6 × 103)]1/2
= 813.14 ≈ 813 m/s
The velocity of the exit/discharge steam is 813 m/s. ... Ans.
Chemical Engineering Thermodynamics - I 4.92 Second Law of Thermodynamics

Example 4.40 :
Dry saturated steam at 3 bar enters an adiabatic nozzle with negligible velocity at a rate of
1 kg/s and leaves as saturated steam at 373 K.
Calculate : (i) the rate of entropy generation in the nozzle and
(ii) the isentropic efficiency of the nozzle.
Data : Steam at 3 bar : H = 2724.7 kJ/kg and S = 6.9909 kJ/(kg·K)
Steam at 373 K : HV = 2676 kJ/kg, HL = 419.06 kJ/kg,
SL = 1.3069 kJ/(kg·K) and SV = 7.3554 kJ/(kg·K)
Solution :
S1 = Entropy at inlet = 6.9909 kJ/(kg·K) ... steam at 3 bar
S2 = Entropy at exit = 7.3554 kJ/(kg·K) ... saturated steam at 373 K
.
m = Steam rate = 1 kg/s
The rate of entropy generation is given by
. .
S = m (S2 – S1)
= 1 × (7.3554 – 6.9909) = 0.3645 kJ/(s·K) ... Ans. (i)
The isentropic efficiency of a nozzle is given by
[u21 / 2]actual
ηN =
[u22 / 2]isentropic
For the actual nozzle,
[u22 / 2]actual = H1 – H2 in kJ/kg as u2/2 is in J/kg
= [2724.7 – 2676] × 103
= 48.7 × 103 J/kg = 48.7 kJ/kg
For an isentropic nozzle,
S2 = S1 i.e., Entropy at exit = Entropy at inlet
S1 = Entropy at inlet = Entropy of steam at 3 bar = 6.9909 kJ/(kg·K)
Let x be the dryness fraction of steam at 373 K.
S1 = 6.9909 = x (SV) + (1 – x) (SL)
6.9909 = x (7.3554) + (1 – x) (1.3069)
∴ x = 0.9397 ≈ 0.94
H'2 = H2 for the isentropic nozzle
= Enthalpy of steam at 373 K for the isentropic nozzle
= xHV + (1 – x)HL = 0.94 (2676) + (1 – 0.94) (419.06)
= 2540.58 kJ/kg
Chemical Engineering Thermodynamics - I 4.93 Second Law of Thermodynamics

[u22 / 2]isentropic = H1 – H2' = [2724.7 – 2540.58] × 103


= 184.12 × 103 J/kg = 184.12 kJ/kg
[u22 / 2]actual 48.7
∴ ηN = 2 = 184.12 = 0.2645 ... Ans. (ii)
[u2 / 2]isentropic
Example 4.41 :
A small experimental turbine is fed with superheated steam at 1 MPa and 473 K at a rate of
1 kg/s. The steam leaves the turbine at 373 K. Determine the power output of the turbine
assuming that the turbine is reversible and adiabatic.
Data : Steam at 373 K = HV = 2676 kJ/kg, HL = 419.06 kJ/s
SL = 1.3069 kJ/(kg·K) and SV = 7.3554 kJ/(kg·K)
Superheated steam at 1 MPa and 473 K : H = 2826.8 kJ/kg, S = 6.6922 kJ/(kg·K)
Solution :
For a reversible and adiabatic turbine, the second law of thermodynamics gives
Entropy at the exit = Entropy at the entrance of the turbine
S2 = S1
Entropy of superheated steam = 6.6922 kJ/(kg·K)
Entropy at the entrance
of the turbine = Entropy of superheated steam
S1 = 6.6922 kJ/(kg·K)
Let x be the dryness fraction of the steam at 373 K.
S1 = S2 ... Entropy at the exit
6.6922 = xSV + (1 – x) SL
6.6922 = x (7.3554) + (1 – x) (1.3069)
∴ x = 0.89
Enthalpy of steam at 373 K = xHV + (1 – x) HL
= 0.89 (2676) + (1 – 0.89) (419.06)
= 2427.7366 = 2427.74 kJ/kg
According to the first law of thermodynamics,
Ws = – ∆H = – (H2 – H1) = H1 – H2
where H2 = Enthalpy of steam at 373 K = 2427.74 kJ/kg
H1 = Enthalpy of superheated steam = 2826.8 kJ/kg
Ws = 2826.8 – 2427.74
= 399.06 kJ/kg
Chemical Engineering Thermodynamics - I 4.94 Second Law of Thermodynamics

·
m = mass flow rate of superheated steam = 1 kg/s
Power output of ·
the turbine = m Ws

= 1 kg/s × 339.06 kJ/kg


= 339.06
≈ 339 kJ/s = 399 kW … Ans.
Example 4.42 :
Steam at 6 bar and 573 K enters a nozzle at a velocity of 30 m/s. The nozzle operates
isentropically. Find the area of cross-section at a point in the nozzle where the pressure is 4 bar,
as a fraction of the inlet area.
Data : Steam at 6 bar and 573 K : H = 3061.6 kJ/kg, S = 7.3724 kJ/(kg·K)
Specific volume = v = 0.4344 m3/kg
Steam at 4 bar and 473 K : H = 2860.5 kJ/kg, S = 7.1706 kJ/(kg·K)
Specific volume = v = 0.5342 m3/kg
Steam at 4 bar and 523 K : H = 2964.2 kJ/kg, S = 7.3789 kJ/(kg·K)
Specific volume = v = 0.5951 m3/kg
Solution :
For an isentropic nozzle,
S2 = S1
Entropy at the exit = Entropy at the inlet ... of the nozzle
S1 = Entropy at the inlet = Entropy of steam at 6 bar and 573 K
= 7.3724 kJ/(kg·K)
S2 = S1
∴ S2 = 7.3724 kJ/(kg·K)
Let T be the temperature of steam at 4 bar with entropy equal to 7.3724 kJ/(kg.K).
T can be obtained by interpolation.
At 4 bar and 473 K : S = 7.1706 kJ/(kg·K)
At 4 bar and 523 K : S = 7.3789 kJ/(kg·K)

S at 473 K + (T – 473)  523 – 473  = 7.3724 = S2


7.3789 – 7.1706

 

7.1706 + (T – 473)  523 – 473  = 7.3724


7.3789 – 7.1706
 
∴ T = 521.4 K
Temperature of steam = 521.4 K
Chemical Engineering Thermodynamics - I 4.95 Second Law of Thermodynamics

Enthalpy of steam at 4 bar and 521.4 K (H2) can be obtained by interpolation.


H at 473 K = 2860.6 kJ/kg
H at 523 K = 2964.2 kJ/kg
v at 473 K = 0.5342 m3/kg
v at 523 K = 0.5951 m3/kg
H2 = Enthalpy of steam at 4 bar and 521.4 K

= 2860.5 +  523 – 473  [521.4 – 473]


2964.2 – 2860.5
 
= 2960.88 kJ/kg
v = Specific volume of steam at 4 bar and 521.4 K
... by interpolation

= 0.5342 +  523 – 473  [521.4 – 473]


0.5951 – 0.5342
 
= 0.5931 m3/kg
The nozzle is adiabatic. Therefore, Q = 0.
The nozzle is horizontal. Therefore, z1 = z2 and ∆z = 0.
No work is delivered in a nozzle. Therefore, Ws = 0.
For this nozzle : Q = 0, Ws = 0, z1 = z2, ∆z = 0, u1 = 30 m/s, u2 = ?
H1 = 3061.6 kJ/kg and H2 = 2960.88 kJ/kg
u2 = Velocity at the exit
u1 = Velocity at the inlet to the nozzle
H1 = 3061.6 kJ/kg = 3061.6 × 103 J/kg
H2 = 2960.88 kJ/kg = 2960.88 × 103 J/kg
With the above conditions, the first law of thermodynamics for a steady flow process
reduces to
2 2
u2 – u1
(H2 – H1) + 2 = 0 ... in J/kg
2 2
u2 u1
2 = H1 – H2 + 2
(30)2
= 3061.6 × 103 – 2960.88 × 103 + 2

= 101170
∴ u2 = [2 × 101170]1/2 = 449.8 m/s
Chemical Engineering Thermodynamics - I 4.96 Second Law of Thermodynamics

Let A1 be the cross-sectional area at the inlet.


Let A2 be the cross-sectional area at a point where P = 4 bar in the nozzle.
At A1 : u1 = 30 m/s and v1 = 0.4344 m3/kg
At A2 : u2 = 449.8 m/s and v2 = 0.5931 m3/kg
Mass flow rate of steam is constant through the nozzle.
. A1u1 A2 u2
∴ m = v = v
1 2

The area of cross-section at a point in the nozzle at which P = 4 bar as a fraction of the inlet
area is
A2 u1 · v2 30 × 0.5931
A1 = u2 · v1 = 449.8 × 0.4344 = 0.09106
i.e., A2 = 9.106% of A1 or 0.09106 A1 … Ans.
Example 4.43 :
An inventor claims to have developed a refrigerator which maintains 263 K in it, which is
kept in a room where the surrounding temperature is 298 K. It has a COP of 8.5. How do you
evaluate his claim ?
Solution :
The COP claimed by the inventor = 8.5
T1 = Temperature of the surroundings = 298 K
T2 = Temperature of the cold space = 263 K
The COP of a refrigerator is given by
T2
COP = T – T
1 2

263
= 298 – 263

= 7.514 … maximum possible COP.


Since the COP claimed by the inventor (8.5) is higher than the maximum possible (7.514) for
the given temperature range, the claim of the inventor is invalid. … Ans.
Example 4.44 :
An inventor makes the following claims :
(a) A flame at 1500 K is used as a heat source and the low temperature reservoir is at
300 K. The inventor states that 69% of the heat transfer to the engine from the flame is
returned as work.
(b) A lecture hall receives heat at a rate of 50000 kJ/h from a heat pump. The inside of the
hall is maintained at 295 K and the surroundings are at 272 K. The inventor claims that
work input at a rate of 7000 kJ/h is sufficient.
Chemical Engineering Thermodynamics - I 4.97 Second Law of Thermodynamics

(c) An engine operates between 1000 K and 400 K. The heat transfer to the engine is
500 kW. The inventor states that the work output of 250 kW and the heat rejected to the
low temperature reservoir is 250 kW. State, whether his claims are valid or not ? Why ?
Solution :
(a) As per the claim of the inventor the work output from the engine is 69% of the heat
received by the engine.
Heat absorbed/received by the engine = Q1
Work output from the engine = W = 0.69 Q1
The efficiency of the actual engine is
W 0.69 Q1
ηactual = Q = Q1 = 0.69 or 69%
1

Source temperature = T1 = 1500 K


Sink temperature = T2 = 300 K
The efficiency of a Carnot engine is given by
T1 – T2
ηCarnot = T
1

1500 – 300
= 1500
ηCarnot = 0.80 or 80%
The Carnot efficiency is the maximum possible efficiency. The actual engine should have an
efficiency less than the Carnot efficiency.
Since ηactual (0.69) < ηCarnot (0.80), the inventor's claim is valid. … Ans. (a)
(b) Q1 = Heat received by the hall = 50000 kJ/h
W = Work output required (minimum)
T1 = Temperature of the hall = 295 K
T2 = Temperature of the surroundings = 272 K
The coefficient of performance of a heat pump is given by
Q1 T1
(COP)HP = W = T – T
1 2

Q1 T1
W = T1 – T2

50000 295
Substituting the values, W = 295 – 272
W = 3898.3 kJ/h
Chemical Engineering Thermodynamics - I 4.98 Second Law of Thermodynamics

The minimum work input required is 3898.3 kJ/h and the inventor claims that the work input
of 7000 kJ/h is sufficient.
Since the minimum work input required is less than that supplied (7000 kJ/h), the claim of
the inventor is valid. … Ans. (b)
(c) Given : W = Work output of the engine = 250 kW
Q1 = Heat absorbed by the engine = 500 kW

The efficiency of the engine (as per the inventor's claim) is


W 250
ηactual = Q = 500 = 0.50 or 50%
1

The efficiency of a Carnot engine is given by


T1 – T2
ηCarnot = T1
where, T1 = Source temperature = 1000 K
T2 = Sink temperature = 400 K
1000 – 400
ηCarnot = 1000 = 0.60 or 60%

Since the efficiency of the actual engine is less than the efficiency of the Carnot engine
(maximum possible efficiency), the claim of the inventor is valid. … Ans. (c)
Example 4.45 :
A reversible heat engine as shown in Fig. 4.45 (a) absorbs 800 kJ as heat from a reservoir at
600 K and rejects 100 kJ as heat to a reservoir at 300 K. Find the work done by the engine and
the heat interaction with a reservoir at 400 K.

Fig. 4.45 (a)


Chemical Engineering Thermodynamics - I 4.99 Second Law of Thermodynamics

Solution :

Fig. 4.45 (b)


The given figure shows that Q3 is absorbed by the engine. We will find whether Q3 is
absorbed from the reservoir at 400 K by the engine or rejected by to the reservoir at 400 K.
The first law of thermodynamics about the engine gives
Q1 + Q3 = W + Q2
With Q1 = 800 kJ and Q2 = 100 kJ
800 + Q3 = W + 100
W = 700 + Q3 … (1)
We have : T1 = 600 K, T2 = 300 K, and T3 = 400 K
For a reversible heat engine, we have
Q1 Q3 Q2
T1 + T3 = T2
800 Q3 100
600 + 400 = 300

Q3 = 400 300 – 600


100 800
 
= 400 [0.3333 – 1.3333]
= – 400 kJ
We have, W = 700 + Q3
= 700 + (–400)
= 300 kJ
Work done by the engine = 300 kJ … Ans.
Heat interaction with the reservoir at 400 K = – 400 kJ … Ans.
The minus sign signifies that the heat is rejected to the reservoir at 400 K.
Heat rejected to the reservoir at 400 K = 400 kJ.
Chemical Engineering Thermodynamics - I 4.100 Second Law of Thermodynamics

Considering the heat rejection to the reservoir at 400 K :

Fig. 4.45 (c)


From the first law, we have
Q1 = Q2 + Q3 + W
800 = 100 + 400 + 300 = 800 … satisfied
For a reversible engine, we have
Q1 Q2 Q3
T1 = T2 + T3
800 400 100
600 = 400 + 300
1.333 = 1 + 0.333 = 1.333 … satisfied
Example 4.46 :
Two mol of an ideal gas occupying volume of 2 dm3 at 300 K are heated to 325 K. If the
volume due to heating becomes 4 dm3, calculate the entropy change of the gas.
Data : Cv = 12.5 J/(mol.K)
Solution :
T1 = 300 K, V1 = 2 dm3 = 0.20 m3 [d ⇒ deci ⇒ 10–1]
T2 = 325 K, V2 = 4 dm3 = 0.40 m3
n = 2 mol, Cv = 12.5 J/(mol.K)
R = 8.31451 J/(mol.K)
The entropy change as a function of T and V is given by
T2 V2
∆S = nCv ln T + n R ln V
1 1

325 0.40
= 2 × 12.5 ln 300 + 2 × 8.31451 ln 0.20
= 2.0 + 11.526
= 13.526 ≈ 13.53 J/K … Ans.
[mol × J/(mol.K) = J/K]
Chemical Engineering Thermodynamics - I 4.101 Second Law of Thermodynamics

Example 4.47 :
One mole of an ideal gas is heated from 100 K to 300 K. Calculate ∆S if (i) the pressure is
kept constant and (ii) the volume is kept constant. Take Cv = 1.5 R.
Solution :
(i) For constant P,
T2
∆S = Cp ln T
1

We know that, Cp – Cv = R, ∴ Cp = Cv + R
T2
∆S = (Cv + R) ln T
1

T2
= (1.5 R + R) ln T
1

T2
= 2.5 R ln T
1

R = 8.31451 J/(mol.K), T2 = 300 K and T1 = 100 K

∆S = 2.5 × 8.31451 ln 100


300
 
= 22.84 J/(K.mol) … Ans.
T2
For constant V, ∆S = Cv ln T
1

T2
= 1.5 R ln T = 1.5 × 8.31451 ln 100
300
1  
= 13.70 J/(K.mol) … Ans.
Example 4.48 :
Calculate the entropy change for melting of 1 kg of ice at 273 K (0 °C). The heat of fusion of
ice is 334.72 J/g.
Solution :
Tf = fusion temperature = 273 K
∆Hfusion = latent heat of fusion = 334.72 J/g = 334.72 kJ/kg
The entropy change of melting of ice is
∆Hfusion
∆Sfusion = Tf
334.72
= 273
= 1.226 kJ/(K.kg)
= 1226 J/(K.kg) … Ans.
Chemical Engineering Thermodynamics - I 4.102 Second Law of Thermodynamics

Example 4.49 :
Calculate the entropy change associated with freezing of 1 mol of water at 298 K to ice at
263 K using the following data :
(i) Heat of fusion of ice at its fusion point (273 K) = 6.00 kJ/mol.
(ii) Cp of ice = 36.82 J/(mol.K)
(iii) Cp of water = 75.31 J/(mol.K)
Solution :
The irreversible freezing of water from 298 K to ice at 263 K can be supposed to consist of
the following three reversible processes :
(1) H2O (l) at 298 K → H2O (l) at 273 K
(2) H2O (l) at 273 K → H2O (s) at 273 K
(3) H2O (s) at 273 K → H2O (s) at 263 K.
The total entropy change is equal to the sum of the entropy changes involved in the three
processes.
∆S = ∆S1 + ∆S2 + ∆S3
(1) For H2O (l) at 298 K to H2O (l) at 273 K :
Tf
dT
⌠ Cp d ln T = Cp ⌠
∆S1 = ⌡ ⌡ T
T1

= Cp ln T 
Tf
 1
where, Cp = Cp of water = 75.31 J/(mol.K)
Tf = 273 K and T = 298 K
273
∆S1 = 75.31 ln 298

= – 6.598 ≈ – 6.6 J/(mol.K)


(2) For H2O (l) at 273 K to H2O (s) at 273 K :
– ∆Hfusion
∆S2 = Tf
[latent heat of freezing = – latent heat of fusion/melting]
∆Hfusion = 6.00 kJ/mol = 6000 J/mol
Tf = 373 K
6000
∆S2 = – 273 = – 21.978 ≈ – 22 J/(mol.K)
Chemical Engineering Thermodynamics - I 4.103 Second Law of Thermodynamics

(3) For H2O (s), i.e., ice at 273 K to H2O (s) at 263 K :
T T2
∆S3 = Cp ln T , from ∆S3 = Cp ln T
f 1

where, Cp = Cp of ice = 36.82 J/(mol.K)


T = 263 K and Tf = 273 K
263
∆S3 = 36.82 ln 273 = – 1.37 J/(mol.K)
The entropy change is
∆S = ∆S1 + ∆S2 + ∆S3
= – 6.6 + (– 22) + (– 1.37) = – 29.97 J/(mol.K) … Ans.
Example 4.50 :
Calculate the entropy change associated with conversion of ice at 273 K into one mole of
steam at 373 K.
Data : Latent heat of fusion at melting point = 333.56 × 103 J/kg
Latent heat of vaporisation at boiling point = 2491 × 103 J/kg
Average Cp of water = 4.184 × 103 J/(kg.K)
Solution :
The process of conversion of ice is considered to consist of the following reversible steps :
(1) H2O (s) at 273 K → H2O (l) at 273 K
(2) H2O (l) at 273 K → H2O (l) at 373 K
(3) H2O (l) at 373 K → H2O (v), i.e., steam at 373 K.
Mol. Wt. of H2O = 18 kg/kmol = 18 × 10–3 kg/mol
Let us convert the data from per kg to per mol basis.
Cp = 4.184 × 103 J/(kg.K) = 4.184 × 103 J/(kg.K) × 18 × 10–3 kg/mol
= 75.31 J/(mol.K)
Latent heat of vaporisation = ∆Hvap = 2491 × 103 J/kg
= 2491 × 103 × 18 × 10–3
= 44838 J/mol
Latent heat of fusion = ∆Hfusion = 333.56 × 103 J/kg
= 333.56 × 103 × 18 × 103
= 6004 J/mol
Tb = 373 K and Tf = 273 K
The total entropy change is the sum of the entropy changes in the three steps.
∆S = ∆S1 + ∆S2 + ∆S3
(1) ∆S1
∆Hfusion 6004
∆S1 = Tf = 273 = 21.99 = 22 J/(mol.K)
Chemical Engineering Thermodynamics - I 4.104 Second Law of Thermodynamics

(2) ∆S2
T2 Tb 373
∆S2 = Cp ln T = Cp ln T = 75.31 × ln 273 = 23.50 J/(mol.K)
1 f
∆Hvap 44838
(3) ∆S3 = Tb = 373 = 120.21 J/(mol.K)
∴ ∆S = ∆S1 + ∆S2 + ∆S3
= 22 + 23.50 + 120.21
∆S = 165.71 J/(mol.K) … Ans.
Example 4.51 :
Calculate the entropy of mixing of one mole of oxygen gas with 2 mol of nitrogen gas,
assuming that the gas mixture behaves ideally.
Solution :
The entropy of mixing is given by
∆Sm = – R ∑ ni ln xi
= – R [n1 ln x1 + n2 ln x2] … for mixing of two gases
Moles of oxygen = n1 = 1 mol
Moles of nitrogen = n2 = 2 mol
Total moles of the gas mixture = n = 1 + 2 = 3 mol
x1 = mole fraction of oxygen in the gas mixture = 1/3 = 0.333
x2 = mole fraction of nitrogen in the gas mixture = 2/3 = 0.667
R = 8.31451 J/(mol.K)
∆Sm = – R [n1 ln x1 + n2 ln x2]
= – 8.31451 [1 × ln (0.333) + 2 ln (0.667)]
= – 8.31451 [– 1.0996 + (– 0.8099)]
= 15.876 J/K … Ans.
Example 4.52 :
One mol of oxygen gas is mixed with three mol of hydrogen gas at 298 K to form a gas
mixture at 1 bar, the pressure of each gas before mixing is also 1 bar. Calculate the molar
entropy of mixing.
Solution :
As we have to find out the molar entropy of mixing, i.e., entropy of mixing per mol of the gas
mixture, we have to use the relation
∆S of mixing = ∆Sm = – R ∑ xi ln xi
∆Sm = – R [x1 ln x1 + x2 ln x2] … for mixing of two gases
Moles of O2 = n1 = 1 mol, moles of H2 = n2 = 3 mol
Moles of gas mixture = n = 1 + 3 = 4 mol
n1 1
x1 = mole fraction of O2 = n = 4 = 0.25
3
x2 = mole fraction of N2 = 4 = 0.75, R = 8.31451 J/(mol.K)
Chemical Engineering Thermodynamics - I 4.105 Second Law of Thermodynamics

∆Sm = – 8.31451 [0.25 ln (0.25) + 0.75 ln (0.75)]


= – 8.31451 [– 0.3466 + (– 0.2158)]
= 4.676 J/(mol.K) … Ans.
Example 4.53 :
Calculate the standard entropy change of the reaction
2 SO2 (g) + O2 (g) → 2 SO3 (g)
Data : So values for SO2, O2 and SO3 are 248.53, 205.03 and 256.23 J/K per mol,
respectively.
Solution :
2 SO2 (g) + O2 (g) → 2 SO3 (g)
o
The standard entropy change (∆SR ) [i.e., ∆S at 298 K] of this reaction is given by
o o o
∆SR = ∑ Sproducts – ∑ Sreactants
o o o
= [2 × SSO3 ] – [2 × SSO2 + 1 × SO2 ]
= [2 × 256.23] – [2 × 248.53 + 1 × 205.03]
= – 189.63 J/K [per 2 mol SO2 reacted] … Ans.
Example 4.54 :
Calculate the standard entropy change of the reaction
N2 (g) + O2 (g) → 2 NO (g)
Data : Component So, J/(mol.K)
NO (g) 210.45
O2 (g) 205.03
N2 (g) 191.62
Solution :
N2 (g) + O2 (g) → 2 NO (g)
o
The standard entropy change (∆SR ) of this reaction is given by
o o o
∆SR = ∑ Sproducts – ∑ Sreactants
o o o
= [2 × SNO ] – [1 × SN2 + 1 × SO2 ]
= [2 × 210.45] – [1 × 191.62 + 1 × 205.03]
= 24.25 J/K [per mol N2 reacted] … Ans.
Example 4.55 :
o
Calculate ∆SR for the reaction
H2 (g) + Cl2 (g) → 2 HCl (g) at 500 K
Data : The standard entropies of H2 (g), Cl2 (g) and HCl (g) are 130.60, 222.96 and
o
186.22 J/(mol.K) respectively. CP data :
H2 (g) : 28.6105 + 1.094 × 10–3 T – 0.1476 × 10–6 T2
Cl2 (g) : 28.5463 + 23.879 × 10–3 T – 21.3631 × 10–6 T2
HCl (g) : 30.3088 – 7.609 × 10–3 T + 13.2608 × 10–6 T2
Chemical Engineering Thermodynamics - I 4.106 Second Law of Thermodynamics

Solution :
H2 (g) + Cl2 (g) → 2 HCl (g)
The standard entropy change for this reaction at 298 K is given by
o o o
∆SR = ∑ Sproducts – ∑ Sreactants
o o o o
∆SR = [2 × SHCl ] – [1 × SH2 + 1 × SCl2 ]
= [2 × 186.22] – [1 × 130.60 + 1 × 222.96]
= 18.88 J/K [per mol H2 reacted]
o
We have to find ∆SR at T = 500 K.
The standard entropy change of a chemical reaction at any temperature T is given by the
relation
T
o o o
∆SRT = ∆SR + ⌠
⌡ ∆Cp d ln T
298
o o o
where, ∆Cp = ∑ Cp products – ∑ Cp reactants
o o o
= [2 × Cp HCl ] – [1 × Cp H2 + 1 × Cp Cl2 ]
o
∆Cp = 2 × [30.3088 – 7.609 × 10–3 T + 13.2608 × 10–6 T2]
– 1 × [28.6105 + 1.094 × 10–3 T – 0.1476 × 10–6 T2]
– 1 × [28.5463 + 23.879 × 10–3 – 21.3631 × 10–6 T2]
= 3.4608 – 40.191 × 10–3 T + 5.0109 × 10–6 T2
T
o o
⌠ [3.4608 – 40.191 × 10–3 T + 5.0109 × 10–6 T2] d ln T
∆SRT = ∆SR + ⌡
298
o o
∆SR = ∆SR (298)
T
o o dT
∆SRT = ∆SR + ⌠
⌡ [3.4608 – 40.191 × 10 T + 5.0109 × 10 T ] T
–3 –6 2

298
T
o
= ∆SR + ⌠ 3.4608 –6 
⌡  T – 40.191 × 10 + 5.0109 × 10 T dT
–3

298
o o
∆SRT = ∆SR + 3.4608 [ln T – ln 298] – 40.191 × 10–3 [T – 298]
5.0109 × 10–6 2
+ 2 [T – (298)2]
o
Substituting T = 500 K and ∆SR = 18.88 J/K per mol H2 gives
o
∆SR 500 = 18.88 + 3.4608 [ln 500 – ln 298] – 40.191 × 10–3 [500 – 298]
5.0109 × 10–6
+ 2 [(500)2 – (298)2]
= 18.88 + 1.7909 – 8.118 + 0.4039
= 12.9568 ≈ 12.96 J/K per mol H2 reacted … Ans.
❐❐❐
Chapter ... 5
THERMODYNAMIC PROPERTIES OF
PURE FLUIDS
Thermodynamic properties of fluids can be classified into three groups as :
(1) Reference properties : Pressure (P), volume (V), temperature (T) and entropy (S).
(2) Energy properties : Internal energy (U), enthalpy (H), Gibbs free energy (G) and
Helmholtz free energy and
(3) Derived properties : Heat capacity (Cp and Cv), coefficient of volume expansion (β),
a Joule-Thomson coefficient (µ) and isothermal compressibility (κ).
The thermodynamic properties mentioned above can be classified as measurable properties -
P, V, T, Cp, Cv, β and κ and non-measurable properties - U, H, S, G and A.
In order to determine the non-measurable properties, it is necessary to relate these properties
to the measurable properties like P, V, T, Cp, etc., which are required for the evaluation of heat
and work requirements for industrial processes. In this chapter, the relationships between the
non-measurable and measurable thermodynamic properties are presented using the partial
derivatives and Jacobian methods.
Since the Gibbs free energy and Helmholtz free energy predict the feasibility of a
process or reaction and equilibrium conditions at constant temperature and pressure or volume
[∆G ≤ 0 (at constant T and P), ∆A ≤ 0 (at constant T and V)], these energy properties involving
entropy in their definition are more convenient for the use than the entropy. Therefore, let us
define the Helmholtz free energy and the Gibb’s free energy.
HELMHOLTZ FREE ENERGY
The Helmholtz free energy of a system is defined as
A = U – TS ... (5.1)
Since the properties U, S and T depend upon the state of the system, it is evident that A also
depends upon the state of the system. Therefore, A is a state function and like U and S, it is an
extensive property.
For an isothermal reversible change from state 1 to state 2, the change in the function A of
the system associated with this process is
A2 – A1 = (U2 – U1) – T (S2 – S1)
i.e., ∆A = ∆U – T ∆S ... (5.2)
where ∆S is the change in entropy and ∆U is the change in internal energy of the system.
(5.1)
Chemical Engineering Thermodynamics - I 5.2 Thermodynamic Properties of Pure Fluids

Since the process is reversible, the heat absorbed at constant T is equal to Qrev, given by
Qrev = T ∆S
With this Equation (5.2) becomes
∆A = ∆U – Qrev ... (5.3)
The first law of thermodynamics for a closed system is
∆U = Q – W , Q = Qrev
Substituting this in Equation (5.3), we get
∆A = Qrev – W – Qrev
∆A = –W
or – ∆A = W ... (5.4)
– ∆A is the decrease in the Helmholtz free energy/work function.
Since the process is carried out reversibly, W represents the maximum work. It is thus clear
from Equation (5.4) that the decrease in the work function (i.e., – ∆A) in an isothermal process is
equal to the maximum work done by the system during the process. Therefore, the decrease in
the work function during an isothermal process is a measure of the maximum work available
from a given change of state. (The Helmholtz free energy i.e. the function A is, therefore, termed
as the work function).
GIBBS FREE ENERGY
The Gibbs free energy is a measure of that portion of the enthalpy in a flow process or
constant-pressure batch process which is available for useful work.
The Gibbs free energy (G) is defined by the equation
G = H – TS ... (5.5)
Energy available = Total energy – Energy unavailable
Like H, S and T, G is a state function. It is an extensive property, which is widely used in the
study of chemical reaction equilibria.
The enthalpy of a system is defined as
H = U + PV
Therefore, Equation (5.5) can be written as
G = (U + PV) – TS
i.e., G = A + PV ... (5.6)
It is clear from Equation (5.6) that A is related to G in the same way as U is related to H.
For an isothermal reversible change occurring from state 1 to state 2, the change in the Gibbs
free energy of the system during this process is given as
(G2 – G1) = (H2 – H1) – T (S2 – S1)
i.e., ∆G = ∆H – T ∆S ... (5.7)
Chemical Engineering Thermodynamics - I 5.3 Thermodynamic Properties of Pure Fluids

The change in enthalpy when the system goes from state 1 to state 2 at constant pressure is
given by
∆H = ∆U + P ∆V ... (5.8)
Substituting Equation (5.8) in Equation (5.7), we get
∆G = ∆U + P ∆V – T ∆S
We have, ∆A = ∆U – T ∆S
Therefore, ∆G = ∆A + P ∆V
Since ∆A = – W [Equation (5.4)]
∆G = – W + P ∆V
or – ∆G = W – P ∆V ... (5.9)
The quantity P ∆V is the work of expansion or displacement work, whereas the quantity W is
the total reversible work that also includes P ∆V.
– ∆G = W – P ∆V = Net work ... (5.10)
The decrease in the Gibbs free energy in a process occurring at constant temperature and
pressure is a measure of the maximum work other than the work of expansion available
(or the maximum net work obtainable) from a given change of state.
[Important : For Preliminary Mathematical Background refer to Chapter 1. This topic
covers general procedures which are to be used for the derivation of thermodynamic property
relations in the treatment to follow.]
FUNDAMENTAL PROPERTY RELATIONS
The first law of thermodynamics for a closed system undergoing a reversible process is given
by (for differential/infinitesimal changes),
dU = dQ – dW ... (5.11)
Since dW = P dV
... reversible work of volume expansion/displacement work
dU = dQ – P dV ... (5.12)
According to the second law of thermodynamics,
dQ = T dS … for reversible process
Therefore, Equation (5.12) becomes
dU = T dS – P dV ... (5.13)
Equation (5.13) is the combined statement of the first and second laws of thermodynamics
for reversible processes.
Since this equation contains only properties of the system which depend on states of the
system and not on the kind of process, it is applicable to any process.
The enthalpy is defined as
H = U + PV
Chemical Engineering Thermodynamics - I 5.4 Thermodynamic Properties of Pure Fluids

Differentiating, we get
dH = dU + P dV + V dP ... (5.14)
Replacing dU by Equation (5.13), this reduces to
dH = T dS – P dV + P dV + V dP
dH = T dS + V dP ... (5.15)
The Gibbs free energy is given by
G = H – TS
Differentiating, we get
dG = dH – S dT – T dS
Replacing dH by Equation (5.15), this equation becomes
dG = S dT + V dP – S dT – T dS
dG = – T dS + V dP ... (5.16)
The Helmholtz free energy is defined as
A = U – TS
Differentiating gives
dA = dU – T dS – S dT
Replacing dU by Equation (5.13), this equation becomes
dA = T dS – P dV – T dS – S dT
dA = – S dT – P dV ... (5.17)
Equations (5.13), (5.15), (5.16) and (5.17) are the fundamental property relations for one
mole of a homogeneous fluid of constant composition (pure substance). These equations express
the energy properties in terms of the reference properties P, V, T and S.
It is clear from Equations (5.13), (5.15), (5.16) and (5.17) that each of U, H, G and A are
functionally related to a special pair of variables.
U = f (S, V)
H = f (S, P)
G = f (T, P)
and A = f (T, V)
Since dU, dH, dG and dA are exact differentials, we can express them as
∂U ∂U
dU =   dS +   dV ... total differential of U ... (5.18)
 ∂S V ∂VS
∂H ∂H
dH =   dS +   dP ... (5.19)
 ∂S P  ∂P S
∂G ∂G
dG =   dT +   dP ... (5.20)
 P
∂T  ∂P T
∂A ∂A
dA =   dT +   dV ... (5.21)
 V
∂T ∂VT
Chemical Engineering Thermodynamics - I 5.5 Thermodynamic Properties of Pure Fluids

MAXWELL’S EQUATIONS
These equations are useful to replace non-measurable quantities appearing in thermodynamic
equations by measurable quantities.
The total differential of U for U = f (S, V) (i.e., U is a function of S and V) is
∂U ∂U
dU =   dS +   dV ... (5.22)
 ∂S V ∂VS
We have, dU = T dS – P dV ... (5.23)
Comparing Equations (5.22) and (5.23), we get
∂U = T
 ∂S  ... (5.24)
 V
∂U = – P
and ∂V ... (5.25)
 S
Differentiating Equation (5.24) with respect to V at constant S gives
 ∂ ∂V  ∂2U  ∂T 
∂V  ∂S   = ∂V ∂S = ∂V ... (5.26)
 V S S

Differentiating Equation (5.25) with respect to S at constant V gives


 ∂ ∂U  ∂2U ∂P
∂S ∂V  = ∂S ∂V = – ∂S ... (5.27)
 S V V

∂2U  ∂T  ∂2U ∂P


=   and = – 
∂V ∂S ∂VS ∂S ∂V ∂SV
Since the order of differentiation of the second mixed partial derivative is immaterial,
one can write
∂2U ∂2U
=
∂V ∂S ∂S ∂V
∂T = – ∂P
Therefore, ∂V  ∂S  ... (5.28)
 S  V
This is the first of the four Maxwell’s equations.
We have, dH = T dS + V dP ... (5.29)
The total differential of H for H = f (S, P) (i.e., H is a function of S and P) is
∂H ∂H
dH =   dS +   dP ... (5.30)
 ∂S P  ∂P S
Comparing Equations (5.29) and (5.30), we get
∂H
 ∂S  = T ... (5.31)
 P
∂H = V
and  ∂P  ... (5.32)
 S
Chemical Engineering Thermodynamics - I 5.6 Thermodynamic Properties of Pure Fluids

Differentiating Equation (5.31) with respect to P at constant S gives


 ∂ ∂H  ∂2H ∂T
∂P  ∂S   = ∂P ∂S = ∂P ... (5.33)
 PS S

Differentiating Equation (5.32) with respect to S at constant P gives


 ∂ ∂H  ∂2H ∂V
∂S  ∂P   = ∂S ∂P =  ∂S  ... (5.34)
 SP P

∂2H ∂2H
Since = ... order of differentiation is immaterial.
∂P ∂S ∂S ∂P
From Equations (5.33) and (5.34), we get
∂T = ∂V
 ∂P    ... (5.35)
 S  ∂S P
The remaining two Maxwell's equations are obtained in a similar way by making use of
Equations (5.16) and (5.20), and Equations (5.17) and (5.21). These equations are as given
below :
 ∂S  =  ∂P 
∂V ∂T ... (5.36)
 T  V
 ∂S  ∂V
–  =   ... (5.37)
 T
∂ P ∂TP
The four Equations (5.28), (5.35), (5.36) and (5.37) are frequently referred to as the
Maxwell’s relations. These equations are applicable for homogeneous as well as heterogeneous
systems at equilibrium.
For the derivation of the Maxwell’s equations/relations and the method of writing the
Maxwell’s relations, please refer to Example 5.59.
CLAPEYRON EQUATION
During a phase change operation from liquid L to vapour V of a pure substance, the
temperature and pressure remain constant. The criterion of equilibrium at constant temperature
and pressure is given by
dG = 0 at constant T and P
Since dG = V dP – S dT, at constant T and P, dT = 0 and dP = 0
and therefore, dG = 0
That is, when both liquid and vapour phases of a pure substance coexist at equilibrium,
the Gibbs free energy of the vapour phase is identical with that of the liquid phase (GL = GV).
Suppose that the temperature of an equilibrium mixture of liquid and vapour is raised by
dT (from T to T + dT), then the pressure must also increase by dPs (from Ps to P + dPs) in
accord with the relation between vapour pressure and temperature for the two phases to coexist at
equilibrium.
Chemical Engineering Thermodynamics - I 5.7 Thermodynamic Properties of Pure Fluids

Let the Gibbs free energy of the substance in the liquid phase at the new temperature, T + dT
and pressure P + dP be GL + dGL and that in the vapour phase be GV + dGV. Since the two phases
are still in equilibrium,
GL + dGL = GV + dGV (since dG = 0)
∴ dGL = dGV (since GL = GV)
The Gibbs free energy in the differential form, for a pure substance is given by
dG = – S dT + V dPs ... (5.38)
For the liquid phase, it becomes
dGL = – SL dT + VL dPs
For the vapour phase, it becomes
dGV = – SV dT + VV dPs
We have : dGL = dGV
– SL dT + VL dPs = – SV dT + VV dPs
dPs SV – SL
dT = VV – VL ... (5.39)

where VV and VL are molar or specific volumes of the saturated vapour and liquid
(m3/mol or m3/kg) respectively.
∆V = VV – VL
= Increase in the volume accompanying the transfer of 1 mol or
1 kg of liquid to the vapour state or the change in the molar
volume due to vaporisation
For the phase change, the change in entropy is given by
HV – HL ∆HV
SV – SL = T = T ... (5.40)

where ∆HV is the molar or specific latent heat of vaporisation/enthalpy of vaporisation


(kJ/mol or kJ/kg).
SV is the molar entropy of the saturated vapour, SL is the molar entropy of the saturated
liquid and SV – SL is the change in the molar entropy due to vaporisation. Ps is the vapour
pressure (saturation pressure) of the liquid at T.
Substituting Equation (5.40) in Equation (5.39) and replacing (VV – VL) = ∆V, we get
dPs ∆ HV
dT = T ∆V ... (5.41)

Equation (5.41) is known as the Clapeyron equation. This equation gives the rate of change
of vapour pressure of the liquid with the temperature.
Chemical Engineering Thermodynamics - I 5.8 Thermodynamic Properties of Pure Fluids

The boiling point of a liquid is the temperature at which the pressure of the vapour in
equilibrium with the liquid is equal to the external pressure P. If the equation is inverted, it gives
the variation of the boiling point of a liquid with the external pressure P.
The Clapeyron equation is used to calculate the latent heat of vaporisation if the variation of
vapour pressure with temperature is known. Alternately, it is used to determine the variation of
vapour pressure with temperature if the latent heat of vaporisation is known.
CLAUSIUS - CLAPEYRON EQUATION
If the phase change from liquid to vapour occurs at very low pressures, then the vapour phase
can be assumed to behave as an ideal gas. Since the volume of the liquid (VL) is very small in
comparison with that of the vapour (VV) at the same P and T, it is reasonable to replace
∆V (= VV – VL) by VV. With this Equation (5.41) becomes
dPs ∆HV
dT = TVV ... (5.42)

For an ideal gas, PV = RT


Therefore, the molar volume VV can be replaced by RT/Ps. With this, Equation (5.42)
becomes
dPs ∆HV
dT = T (RT/P)
Rearranging gives
1 dPs d ln Ps ∆ HV
... since x = d ln x ... (5.43)
dx
= =
s
P dT dT RT2  
Equation (5.43) is known as the Clausius - Clapeyron equation.
If the latent heat of vaporisation (∆HV) is assumed constant over a range of temperatures,
general integration of Equation (5.43) gives
∆HV dT

⌡ d ln Ps = ⌠
R ⌡ T2

∆HV 1
ln Ps = – T T + constant ... (5.44)
 
Equation (5.44) is the equation of a straight line. A plot of d ln Ps versus 1/T yields a straight
line with a slope of – ∆HV/R.
Integration of Equation (5.43) between the temperature limits of T1 and T2 and the
s s
corresponding vapour pressures P1 and P2 , for constant ∆HV in the range T1 to T2 , yields

 Ps2 ∆HV 1 1
ln  s = R T – T  ... (5.45)
P   1 2
 1
Chemical Engineering Thermodynamics - I 5.9 Thermodynamic Properties of Pure Fluids

which is the integrated form of the Clausius - Clapeyron equation. This equation is used
(i) to calculate the latent heat of vaporisation of a liquid if we know the vapour pressures
of the liquid at two different temperatures.
(ii) to calculate the vapour pressure at one temperature if we know the latent heat of
vaporisation and the vapour pressure at another temperature.
(iii) to calculate the boiling point at a given pressure if we know the latent heat of
vaporisation and the boiling point at another pressure (in this Ps is to be replaced by P).
VAPOUR PRESSURE - TEMPERATURE RELATIONSHIPS
The vapour pressure of a liquid at a given temperature is defined as the pressure exerted by
the vapour in equilibrium with the liquid at that temperature.
The Clausius - Clapeyron equation in the general integrated form is given by
∆HV 1
ln Ps = – R T + constant
 
So, according to this equation, the vapour pressure of a liquid can be expressed as
B
ln Ps = A – T ... (5.46)

where A and B are constants. It gives an approximate value of the vapour pressure.
The Antoine equation which is more satisfactory for the estimation of vapour pressure is
given by
B
ln Ps = A – T + C ... (5.47)

where A, B and C are constants. In this equation, Ps is in Torr (mm Hg) and T is in °C.
VOLUME EXPANSIVITY AND ISOTHERMAL COMPRESSIBILITY
Gases and liquids expand on heating, but the expansion of gases is much more than that of
liquids.
The increase in volume of a fluid per unit volume for unit increase of temperature at
constant pressure is called the coefficient of volume expansion or volume expansivity of the fluid.
It is denoted by the symbol β and has the dimension of T–1.
The coefficient of volume expansion / the volume expansivity is defined as
1 ∂V
β = V  ... (5.48)
 ∂T P
Similarly, fluids can be compressed. The decrease in volume of a fluid per unit volume for
unit increase in pressure at constant temperature is called the isothermal compressibility or
coefficient of isothermal compressibility of the fluid. It is denoted by the symbol κ and has the
dimensions of P–1.
Chemical Engineering Thermodynamics - I 5.10 Thermodynamic Properties of Pure Fluids

The isothermal compressibility is defined as


1 ∂V
κ = –V  ... (5.49)
 ∂P T
Both β and κ are measurable properties.
∂V
The negative sign is included in the definition to make κ positive since   is always
 ∂P T
negative (since volume decreases as pressure increases).
For an ideal gas, PV = RT
RT
i.e. V = P

Differentiating with respect to T at constant P gives


∂V = R
 ∂T 
 P P

β = V  P  = PV = T since PV = T
1 R R 1 R 1

 
Differentiating V = RT/P with respect to P at constant T gives
∂V = – RT
 ∂P 
 T P2

κ = – V  P2  = – PV × P = – PV × P = P
1 RT RT 1 PV 1 1

     
1 1
So, for an ideal gas, β = T and κ = P
ENTROPY - HEAT CAPACITY RELATIONSHIPS
The heat capacity at constant volume of a substance is defined as
∂Q
Cv =   ... (5.50)
 ∂T V
We have, dU = dQ – P dV
dV = 0, for a constant volume process. Therefore, dQ = dU
Hence, the heat capacity at constant volume in terms of the property (U) change of the
system is defined as
∂U
Cv =   ... (5.51)
 ∂T V
We know that dQ = T dS for a reversible process. Therefore, Equation (5.50) can be
written as
∂S
Cv = T   ... (5.52)
∂TV
Chemical Engineering Thermodynamics - I 5.11 Thermodynamic Properties of Pure Fluids

Rearranging gives
∂S = Cp
∂T ... (5.53)
 V T
The heat capacity at constant pressure of a substance is defined as
∂Q
Cp =   ... (5.54)
 ∂T P
We know that dU = dQ – P dV
For a constant pressure process,
P dV = d (PV)
Therefore, dU = dQ – d (PV)
dQ = dU + d (PV) = d (U + PV) = dH
Hence, the heat capacity at constant pressure in terms of the property (H) change of the
system is defined as
∂H
Cp =   ... (5.55)
 ∂T P
For a reversible process, we know that dQ = T dS. Therefore, Equation (5.54) can be written
as
∂S
Cp = T   ... (5.56)
∂TP
On rearrangement it becomes
∂S = Cp
∂T ... (5.57)
 P T
The above given equations are used to determine the entropy change of a substance at
constant pressure or at constant volume for a specified change in temperature.
Integration of Equation (5.57) gives

∆S = ⌠
⌡ Cp dT
If Cp is independent of temperature, then
T2
∆S = Cp ln T ... (5.58)
1

DIFFERENCE BETWEEN HEAT CAPACITIES


The first T dS equation is
∂P
T dS = Cv dT + T   dV ... Equation (5.87)
∂TV
The second T dS equation is
∂V
T dS = Cp dT – T   dP ... Equation (5.98)
 ∂T P
Chemical Engineering Thermodynamics - I 5.12 Thermodynamic Properties of Pure Fluids

Equating the first and second T dS equations, we get


∂P ∂V
T dS = Cv dT + T   dV = Cp dT – T   dP
 V
∂T  ∂T P
∂P ∂V
∴ (Cp – Cv) dT = T   dV + T   dP
∂TV  ∂T P
Rearranging gives
∂P
T 
∂V
T 
∂TV  ∂T P
dT = C – C dV + C – C dP ... (5.59)
p v p v
If T is a function of V and P, then the total differential of T is
 ∂T  ∂T
dT =   dV +   dP ... (5.60)
 P
∂V ∂PV
Comparing Equation (5.59) with Equation (5.60), we get
∂P
T 
∂TV  ∂T 
Cp – Cv = ∂VP ... (5.61)

∂V
T 
 ∂T P ∂T
and Cp – Cv = ∂P V ... (5.62)
From Equation (5.61),
∂P
T 
∂TV ∂P ∂V
Cp – Cv = = T   ∂T 
 
∂T ∂TV  P
∂V
 P
From Equation (5.62),
∂V
T 
 ∂T P ∂P ∂V , ∂x = 1
Cp – Cv = = T   ∂T  ∂y
∂T ∂TV  P  z ∂y
∂P ∂x
 V  z
That is, Equations (5.61) and (5.62) give
∂P ∂V
Cp – Cv = T     ... (5.63)
∂TV ∂TP
Among the variables P, T and V, the cyclic relation is
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P 
 V  P  T
∂P = –1 ∂V  ∂P 
∴ ∂T = –  ∂V
 V  ∂T  ∂V  ∂T P  T
∂V  ∂P 
 P  T
Chemical Engineering Thermodynamics - I 5.13 Thermodynamic Properties of Pure Fluids

∂P
Substituting for   by the above value, Equation (5.63) becomes
∂TV
2
∂V  ∂P 
Cp – Cv = – T     ... (5.64)
∂TP ∂VT
1 ∂V 1 ∂V
We know that β = V   and κ = –V  
 ∂T P  ∂P T
∂V = Vβ 1 1
 ∂T  and = –
 P  
∂V κV
 ∂P 
 T
 ∂P  = – 1
∴ ∂V
 T κV
Substituting for the partial derivatives, Equation (5.64) becomes
Cp – Cv = – T (Vβ)2 –
1
 κV
TV β 2
Cp – Cv = … (5.65)
κ
(i) Let us apply Equation (5.64) to a substance following the ideal gas equation.
For an ideal gas, PV = RT
RT
∴ V = P
Differentiating w.r.t. T at constant P yields
∂V = R
 ∂T  ... (5.66 A)
 P P
RT
P = V
Differentiating w.r.t. V at constant T yields
 ∂P  = – RT
∂V ... (5.66 B)
 T V2
Substituting Equations (5.66 A) and (5.66 B) in Equation (5.64), we get
Cp – Cv = – T  P2  – V2  = P2 V2
R2 RT R3 T2
   
(RT) 2 (RT)2
= R (PV)2 = R · (RT)2 = R
∴ Cp – Cv = R ... (5.67)
2
∂V  ∂P 
(ii) Since   is always positive and   is negative, Cp – Cv is always positive.
 ∂T P ∂VT
Hence, Cp is always greater than Cv.
(iii) As T → 0 K, Cp → Cv or at absolute zero, Cp = Cv.
Chemical Engineering Thermodynamics - I 5.14 Thermodynamic Properties of Pure Fluids

RATIO OF HEAT CAPACITIES


∂S ∂S
We have, Cp = T   and Cv = T  
∂TP ∂TV
∂S
∂T
Cp  P
∴ Cv = ∂S  ... (5.68)
∂T
 V
If S is a function of P and T, then the total differential of S is
∂S ∂S
dS =   dP +   dT
∂PT ∂TP
For constant entropy, dS = 0. Therefore, the above equation becomes
∂S ∂S
0 =   dPS +   dTS
 T
∂P ∂TP
This can be written as
∂S = – ∂S ∂P
∂T ∂P ∂T ... (5.69)
 P  T  S
Let S = f (V, T). The total differential of S is
 ∂S  ∂S
dS =   dV +   dT
 T
∂V ∂TV
For constant entropy, dS = 0. Therefore, the above equation becomes
 ∂S  ∂S
0 =   dVS +   dTS
∂VT ∂TV
Rearranging, we get
∂S = –  ∂S  ∂V
∂T ∂V  ∂T  ... (5.70)
 V  T  S
Substituting Equations (5.69) and (5.70) in Equation (5.68), we get
∂S · ∂P
∂P ∂T
Cp  T  S
= ... (5.71)
Cv  ∂S  · ∂V
∂V  ∂T 
 T  S
The chain rule of partial differentiation is
∂z = ∂z · ∂x
∂a   ∂a 
 y ∂xy  y
 ∂P  = ∂P ·  ∂T 
∴ ∂V   ∂V ... (5.72)
 S ∂TS  S
Chemical Engineering Thermodynamics - I 5.15 Thermodynamic Properties of Pure Fluids

Equation (5.71) can be written as


∂S · ∂P ·  ∂T 
∂P ∂T ∂V
Cp  T  S  S
Cv = ... (5.73)
 ∂S 
∂V
 T
Combining Equations (5.72) and (5.73), we get
∂S ·  ∂P 
∂P ∂V
Cp  T  S
= ... (5.74)
Cv  ∂S 
∂V
 T
 ∂S  = ∂P
We have ∂V  
 T ∂TV
∂S = – ∂V
and ∂P  ∂T  ... Maxwell’s equation
 T  P
With these relations, Equation (5.74) becomes
∂V  ∂P 
–  ·  
Cp  ∂T P ∂VS ∂V ∂T  ∂P 
Cv = = –  ∂P ∂V ... (5.75)
∂P  ∂T P  V  S
∂T
 V
The cyclic relation among P, V and T is
∂V ∂T  ∂P  = – 1
 ∂T  ∂P ∂V
 P  V  T
∂V ∂T = –1 = – ∂V
∴  ∂T  ∂P  ∂P  ... (5.76)
 P  V  ∂P   T
∂V
 T
Substituting Equation (5.76) in Equation (5.75), we get
Cp  ∂P   ∂V  ∂V  ∂P 
Cv = – ∂VS × –  ∂P T =  ∂P T · ∂VS ... (5.77)

Equation (5.77) can be written as

∂V
 ∂P 
Cp  T
Cv = ∂V ... (5.78)
 ∂P 
 S
Chemical Engineering Thermodynamics - I 5.16 Thermodynamic Properties of Pure Fluids

The adiabatic compressibility is defined as


1 ∂V
κs = – V  
 ∂P S
∂V = – V κ
∴  ∂P  s ... (5.79)
 S
1 ∂V
We know that κ = –V 
 ∂P T
∂V = – V κ
∴  ∂P  ... (5.80)
 T
Substituting Equations (5.79) and (5.80) in Equation (5.78), we get
Cp – Vκ
Cv = –V κs = γ
Cp κ
∴ Cv = γ = κs ... (5.81)

DIFFERENTIAL EQUATIONS FOR ENTROPY


The entropy of a substance can be expressed as a function of T and P or T and V, may be
expressed mathematically as
S = f (T, P) and S = f (T, V)
If S is a function of T and V, then the total differential of S is
∂S
dS =   dT +   dV
 ∂S  ... (5.82)
 V
∂T ∂VT
The heat capacity at constant volume can be defined as
∂S
Cv = T  
∂TV
Rearranging gives
∂S = Cv
∂T ... (5.83)
 V T
The Maxwell’s equation is
 ∂S  = ∂P
∂V   ... (5.84)
 T ∂TV
Substituting Equations (5.83) and (5.84) in Equation (5.82), we get
Cv  ∂P 
dS = T dT +   dV ... (5.85)
∂TV
Equation (5.85) is the desired equation for dS in terms of the measurable properties P, V, T
and Cv to estimate the change in entropy of any substance.
This is also applicable to an ideal gas.
Chemical Engineering Thermodynamics - I 5.17 Thermodynamic Properties of Pure Fluids

For an ideal gas, PV = RT


RT
P = V

∂P = R
∴ ∂T
 V V
With this, Equation (5.85) becomes
Cv dV
dS = T dT + R V ... (5.86)
Multiplying Equation (5.85) throughout by T, we get

T dS = Cv dT + T   dV
∂P
... (5.87)
∂TV
This equation is known as the first T dS equation.
∂P
The partial derivative   appearing in Equation (5.87) can be expressed in terms of
∂TV
∂V and ∂V . These partial derivatives are respectively related to the coefficients of
 ∂T   ∂P 
 P  T
volume expansion and isothermal compressibility of the substance.
Consider that the volume is a function of P and T.
V = f (P, T)
The total differential of V is
∂V ∂V
dV =   dP +   dT ... (5.88)
 T
∂P  ∂T P
At constant volume, dV = 0. Therefore,
∂V ∂V
0 =   dP +   dT … at constant V
 ∂P T  ∂T P
∂V dP = – ∂V dT
 ∂P   ∂T  ... at constant V
 T  P
∂V
– 
∂P =  ∂T P
∴ ∂T ... (5.89)
 V ∂V
 ∂P 
 T
∂P
Substituting for   from Equation (5.89), Equation (5.85) becomes
∂TV
∂V
 ∂T 
Cv  P
dS = T dT – dV ... (5.90)
∂V
 ∂P 
 T
Chemical Engineering Thermodynamics - I 5.18 Thermodynamic Properties of Pure Fluids

1 ∂V
We have, β = V 
 ∂T P
∂V = V β
∴  ∂T  ... (5.91)
 P
1 ∂V
κ = –V 
 ∂P T
∂V = – V κ
∴  ∂P  ... (5.92)
 T
Combining Equations (5.90), (5.91) and (5.92), we get
Cv β
dS = T dT + dV
κ
Tββ
or T dS = Cv dT + dV ... (5.93)
κ
Consider that the entropy is a function of pressure and temperature, then S = f (P, T) and the
total differential of S is
∂S ∂S
dS =   dP +   dT ... (5.94)
∂PT ∂TP
The heat capacity at constant pressure can be defined as
∂S
Cp = T  
∂TP
∂S = Cp
Rearranging, ∂T ... (5.95)
 P T
∂S = – ∂V
We have, ∂P  ∂T  ... Maxwell’s equation ... (5.96)
 T  P
Substituting Equations (5.95) and (5.96) in Equation (5.94), we get
∂V Cp
dS = –   dP + T dT
 ∂T P
Cp ∂V
i.e., dS = T dT –   dP … (5.97)
∂TP
This equation is used to estimate the change in entropy of a substance. It is the desired
equation of dS in terms of the measurable properties P, V, T and Cp.
Multiplying Equation (5.96) by T throughout, we get

T dS = Cp dT – T   dP
∂V
... (5.98)
∂TP
This equation is known as the second T dS equation. We know that
1 ∂V
β = V 
 ∂T P
Chemical Engineering Thermodynamics - I 5.19 Thermodynamic Properties of Pure Fluids

Therefore, Equation (5.96) can be rewritten as


dT
dS = Cp T – βV dP
or T dS = Cp dT – β TV dP ... (5.99)
Using this equation one can estimate the change in entropy of a substance since this equation
expresses dS in terms of the measurable properties P, V, T, Cp and β.
EFFECT OF P AND V ON CP

We know that
∂S
Cp = T  
∂TP
∂S = Cp
∴ ∂T ... (5.100)
 P T
Differentiating Equation (5.100) with respect to P keeping T constant yields
 ∂ ∂S  ∂2S ∂Cp/T 1 ∂Cp
∂P ∂T  = ∂P ∂T =  ∂P  = T  ∂P  ... (5.101)
 P T T T

∂S = – ∂V
We have : ∂P  ∂T  ... Maxwell’s equation
 T  P
Differentiating this relation with respect to T keeping P constant gives
∂2S ∂2V
= –  2 ... (5.102)
∂T ∂P  ∂T P
∂2S ∂2S
We know that = ... Order of differential is immaterial
∂P ∂T ∂T ∂P
Therefore, from Equations (5.101) and (5.102), we get
1 ∂Cp ∂2V

T  ∂P T = –  ∂T2 
 P
∂Cp
 = – T  2 
∂ 2V
i.e.,  ... (5.103)
 ∂P T ∂T P
Equation (5.103) represents the effect of pressure on Cp.
∂Cp ∂Cp ∂V
We know that :   =     ... Chain rule of partial differentiation
 T  ∂V T  ∂P T
∂P
Therefore, with this, Equation (5.103) becomes
∂Cp ∂V
    = – T  2 
∂2V
 ∂V T  ∂P T  ∂T 
∂Cp
 = – T  2   
∂ 2V ∂P
i.e.,  ... (5.104)
 ∂V T ∂T P ∂VT
1  ∂P 
Since =  
∂V ∂VT
 ∂P 
 T
Equation (5.104) represents the effect of volume on Cp.
Chemical Engineering Thermodynamics - I 5.20 Thermodynamic Properties of Pure Fluids

DIFFERENTIAL EQUATIONS FOR INTERNAL ENERGY :


Let the internal energy U be expressed as a function of T and V.
U = f (T, V). Then the total differential of U is
∂U ∂U
dU =   dT +   dV ... (5.105)
 ∂T V ∂VT
The heat capacity at constant volume is given by
∂U
Cv =   ... (5.106)
 ∂T V
Substituting Equation (5.106) in Equation (5.105), we get
∂U
dU = Cv dT +   dV ... (5.107)
∂VT
We know that dU = T dS – P dV
∂U = T  ∂S  – P
∴ ∂V ∂V ... (5.108)
 T  T
 ∂S  = ∂P
We have : ∂V   ... Maxwell’s equation
 T ∂TV
With this, Equation (5.108) becomes
∂U = T ∂P – P
∂V ∂T ... (5.109)
 T  V
Substituting Equation (5.109) in Equation (5.107) yields

dU = Cv dT + T   – P dV
∂P
... (5.110)
 ∂TV 
This is the required relation for dU in terms of the measurable properties P, V, T and Cv for
estimating the change in the internal energy of a substance.
Let us apply Equation (5.110) to a substance which follows the ideal gas law equation.
PV = RT
RT
P = V
Differentiating this with respect to T at constant V gives
∂P =  ∂ RT = R
∂T ∂T  V 
 V  V V
Substituting this in Equation (5.110),
dU = Cv dT + T · V – P dV
R
 
= Cv dT +  V – P dV
RT
 
RT
dU = Cv dT + [P – P] dV = Cv dT, since P = V
Chemical Engineering Thermodynamics - I 5.21 Thermodynamic Properties of Pure Fluids

The cyclic relation among P, V and T is


∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P 
 V  P  T
∂V
 ∂T 
∂P = – 1  P
∂T = – ... (5.111)
 V  ∂T  ∂V ∂V
∂V  ∂P   ∂P 
 T  T  T
∂P
Substituting for   from Equation (5.111), Equation (5.110) becomes
∂TV
∂P
dU = Cv dT + T   dV – P dV
∂TV
∂V
 ∂T 
 P
dU = Cv dT – T dV – P dV
∂V
 ∂P 
 T
∂V
  ∂T  

dU = Cv dT – P + T
 P 
dV ... (5.112)
 ∂V 
 ∂P 
  T 
1 ∂V
β is defined as β = V 
 ∂T P
∂V = β V
Rearranging,  ∂T 
 P
1 ∂V
κ is defined as κ = –V 
 ∂P T
∂V = – κ V
Rearranging,  ∂P 
 T
∂V
 ∂T 
 P – βV β
∴ = = – ... (5.113)
∂V κV κ
 ∂P 
 T
Combining Equations (5.112) and (5.113), we get

dU = Cv dT – P –
 Tβ 
 dV
 κ
Tβ – P dV
dU = Cv dT + 
i.e.,  ... (5.114)
κ 
We know that : dU = T dS – P dV ... (5.115)
Chemical Engineering Thermodynamics - I 5.22 Thermodynamic Properties of Pure Fluids

For S = f (P, T), the total differential of S is


∂S ∂S
dS =   dP +   dT ... (5.116)
∂PT ∂TP
∂S ∂S = Cp
We know that : Cp = T   , i.e., ∂T
∂TP  P T

∂S = – ∂V … Maxwell's equation


and ∂P  ∂T 
 T  P
∂S ∂S
Substituting for   and   , Equation (5.116) becomes
∂TP ∂PT
∂V Cp
dS = –   dP + T dT ... (5.117)
 ∂T P
Substituting Equation (5.117) in Equation (5.115), we get
Cp ∂V 
dU = T  T dT –   dP – P dV
  ∂T P 
This can be written as

dU = Cp dT – P dV + T   dP
∂V
... (5.118)
  ∂TP 
DIFFERENTIAL EQUATIONS FOR ENTHALPY
Suppose it is desired to estimate the change in the enthalpy of a substance if its temperature
and pressure are simultaneously changed.
Consider H = f (T, P). The total differential of H is
∂H ∂H
dH =   dT +   dP ... (5.119)
 P
∂T  ∂P T
The heat capacity at constant pressure is given by
∂H
Cp =  
 ∂T P
With this, Equation (5.119) becomes
∂H
dH = Cp dT +   dP ... (5.120)
 ∂P T
We know that dH = T dS + V dP
∂H = T ∂S + V
∴  ∂P  ∂P ... (5.121)
 T  T
∂S = – ∂V
We have, ∂P  ∂T  ... Maxwell’s equation
 T  P
Chemical Engineering Thermodynamics - I 5.23 Thermodynamic Properties of Pure Fluids

Substituting this in Equation (5.121), we get


∂H = – T ∂V + V
 ∂P   ∂T  ... (5.122)
 T  P
Substituting Equation (5.122) in Equation (5.120), we get

dH = Cp dT + V – T    dP
∂V
... (5.123)
  ∂TP
This is the desired relationship for dH in terms of the measurable properties P, V, T and Cp.
It is used for estimating the change in enthalpy of a substance.
Let us obtain dH in terms of dV and dT.
∂V
 ∂T 
Cv  P
We have, dS = T dT – dV
∂V
 ∂P 
 T
and dH = T dS + V dP
Substituting for dS by the value given above, the equation for dH becomes
∂V
 ∂T 
 P
dH = Cv dT – T dV + V dP
∂V
 ∂P 
 T
∂V
  ∂T  
dH = Cv dT +
V dP – T  P
dV
 ... (5.124)
 ∂V
 ∂P  
  T 
EFFECT OF TEMPERATURE, PRESSURE AND VOLUME ON U, H AND S
The effect of pressure on entropy is given by
∂S = – ∂V ... Maxwell’s equation
∂P  ∂T  ... (5.125)
 T  P
The effect of temperature on entropy is given by
∂S = Cp
∂T
 P T
The effect of temperature at constant volume on internal energy is given by
dU = Cv dT ... (5.126)
[Equation (5.114) at constant volume (since at constant V, dV = 0) reduces to the above
equation.]
Chemical Engineering Thermodynamics - I 5.24 Thermodynamic Properties of Pure Fluids

The effect of volume on internal energy at constant temperature is given by


∂V
  ∂T  
– P + T 
 P
dU = dV ... (5.127)
 ∂V
 ∂P  
  T 
[Equation (5.118) at constant temperature (since at constant T, dT = 0) reduces to the above
equation.]
The effect of temperature on enthalpy at constant pressure is given by
∂H = C
 ∂T  p ... (5.128)
 P
The effect of pressure on enthalpy at constant temperature is given by
 ∂V 
dH = V – T    dP
 ∂T  P
∂H = V – T ∂V
i.e.,  ∂P   ∂T  ... (5.129)
 T  P
[Equation (5.124) at constant temperature (since at constant T, dT = 0) reduces to the above
equation.]
JOULE - THOMSON COEFFICIENT
The Joule - Thomson coefficient (µ) is defined as the fall in temperature per unit decrease in
pressure at constant enthalpy.
∂T
µ =   ... (5.130)
∂PH
We know that the change in H simultaneously with T and P is given by
 ∂V 
dH = Cp dT + V – T    dP ... (5.131)
∂T
  P
For Joule - Thomson expansion, dH = 0. With this condition, the above equation becomes
 ∂V 
0 = Cp dT + V – T    dP
  ∂T P
Rearranging gives
 ∂V 
Cp dT = T   – V dP
∂T  P 
 ∂V 
T  ∂T  – V
dT =   dP ... at constant H
P

 C p 
Chemical Engineering Thermodynamics - I 5.25 Thermodynamic Properties of Pure Fluids

∂V
T  –V
∂T =  ∂T P
∴ ∂P
 H Cp

∂V
T  –V
 ∂T P
i.e., µ = Cp ... (5.132)

This equation expresses µ in terms of the measurable properties P, V, T and Cp and used for
estimating µ.
GIBBS - HELMHOLTZ EQUATION
We have, dG = – S dT + V dP
At constant P, the above equation becomes
dG = – S dT
∂G = – S
i.e.,  ∂T  ... (5.133)
 P
The derivative G/T with respect to T at constant P is
∂G
T  –G
∂(G/T) =  ∂T P
 ∂T 
 P T2

Substituting Equation (5.133) in the above equation, we get


∂(G/T) = – TS – G = – (G + TS)
 ∂T  ... (5.134)
 P T2 T2

We know that G = H – TS
Therefore, G + TS = H
With this, Equation (5.134) becomes
∂(G/T) = –H
 ∂T  ... (5.135)
 P T2
This is known as the Gibbs - Helmholtz equation.
JACOBIAN METHOD
The thermodynamic relations presented earlier can also be derived by the Jacobian method.
FUNDAMENTAL PROPERTY RELATIONS
We have, dU = T dS – P dV
This can be expressed in terms of Jacobians as
[U, w] = T [S, w] – P [V, w] ... (5.136)
Similarly, dH = T dS + V dP
Chemical Engineering Thermodynamics - I 5.26 Thermodynamic Properties of Pure Fluids

In the Jacobian notation, it becomes


[H, w] = T [S, w] + V [P, w] ... (5.137)
dG = – S dT + V dP
In the Jacobian notation, it is given by
[G, w] = – S [T, w] + V [P, w] ... (5.138)
and dA = – S dT – P dV
In the Jacobian notation, it becomes
[A, w] = – S [T, w] – P [V, w] ... (5.139)
where w is any dummy variable.
MAXWELL’S EQUATIONS
In the Jacobian notation, all the four Maxwell’s equations are given by
[T, S] = [P, V] ... (5.140)
For example, consider the following Maxwell’s equation :
∂P =  ∂S 
∂T ∂V ... (5.141)
 V  T
∂y = [y‚ z]
We know that ∂x
 z [x‚ z]
Therefore, Equation (5.141), in the Jacobian notation, becomes
[P‚ V] [S‚ T]
[T‚ V] = [V‚ T]
It can be written as
[P‚ V] [S‚ T]
[T‚ V] = – [T‚ V] since [x, y] = – [y, x]
Therefore, [P, V] = – [S, T]
[P, V] = – (– [T‚ S])
i.e., [P, V] = [T, S] ... Equation (5.140)
• Cp AND Cv
The heat capacity at constant P, Cp is given by
∂S
Cp = T  
∂TP
In the Jacobian notation, it becomes
[S‚ P]
Cp = T [T‚ P] ... (5.142)
The heat capacity at constant V can be given by
∂S
Cv = T  
∂TV
In terms of the Jacobians, it becomes
[S‚ V]
Cv = T [T‚ V] ... (5.143)
Chemical Engineering Thermodynamics - I 5.27 Thermodynamic Properties of Pure Fluids

• β AND κ
The coefficient of volume expansion is given by
1 ∂V
β = V 
 ∂T P
In the Jacobian notation, it is given by
1 [V‚ P]
β = V [T‚ P] ... (5.144)
The isothermal compressibility is given by
1 ∂V
κ = –V 
 ∂P T
Replacing the partial derivative in favour of the Jacobian,
1 [V‚ T] 1 [V‚ T]
κ = – V [P‚ T] = V [T‚ P] ... (5.145)
FUGACITY
The change in the Gibbs free energy of a pure substance is given by
dG = – S dT + V dP ... (5.146)
At constant T, this equation reduces to
dG = V dP ... (5.147)
If the substance in question is an ideal gas, then
PV = RT
RT
V = P
Substituting this in Equation (5.147), we get
dP
dG = RT P
RT
or dG = P dP = RT d ln P ... (5.148)
Integration of Equation (5.148) yields the simple relation
P2
∆G = RT ln P ... (5.149)
1

Equation (5.149) is valid for ideal gases and is not valid for real gases since for real gases V
RT
is not exactly equal to P .
In order to make this simple free energy equation applicable for real gases, G.N. Lewis
introduced a new function f, called the fugacity. It takes the place of P in Equation (5.148).
Therefore,
dG = V dP = RT ln f ... (5.150)
f2
∆G = RT ln f ... in integrated form ... (5.151)
1
Chemical Engineering Thermodynamics - I 5.28 Thermodynamic Properties of Pure Fluids

Equation (5.151) gives the Gibbs free energy change of any gas (ideal or real) undergoing an
isothermal process between any two states 1 and 2 in terms of fugacity. That is, Equation (5.151)
is valid for real as well as ideal gases.
Fugacity is a kind of fictitious pressure, which is used to retain for real gases, simple forms
of equations which are applicable for ideal gases.
Subtracting Equation (5.148) from Equation (5.150), we get
RT
RT d ln f – RT d ln P = V dP – P dP
RT
∴ RT dln f/P = V dP – P dP
Integrating,
P

RT ln (f/P) = ⌠  RT
... (5.152)
⌡ V – P  dP
P=0
RT
For an ideal gas, V = P and hence from Equation (5.152), f/P = 1.

As P → 0, the behaviour of all gases approaches that of an ideal gas. Therefore, the ratio
f/P approaches unity as P approaches zero, since in that case, real gases approximate to ideal
behaviour. The fugacity, therefore, may be mathematically defined as
lim f f
P→0 P
= 1 or P → 1 as P → 1 ... (5.153)
i.e., as the pressure approaches zero, fugacity approaches pressure. This signifies that the
fugacity is a measure of the pressure of real gases. For an ideal gas, the fugacity is equal to the
pressure. It has dimensions of pressure.
FUGACITY COEFFICIENT
The ratio of the fugacity to the pressure is defined as the fugacity coefficient. It is denoted by
the symbol φ.
f
φ = P ... (5.154)
ACTIVITY
Activity of a substance at any given temperature is defined as the ratio of the fugacity of the
substance to its fugacity in the standard state. It is denoted by the symbol ‘a’.
f
a = fo ... (5.155)
The temperature in the standard state is the same as the temperature in the given conditions.
In the standard state, activity is unity, a = 1 (fo/fo).
The change in the Gibbs free energy of a substance when it undergoes an isothermal process
between the standard state to the given condition is given by
f
∆G = RT ln fo = RT ln a ... (5.156)
We know that dG = V dP
Chemical Engineering Thermodynamics - I 5.29 Thermodynamic Properties of Pure Fluids

If the substance is incompressible between the standard state pressure Po and the given
pressure P, then integration of the above equation yields
∆G = V [P – Po] ... (5.157)
The assumption of constant V holds good in case of solids and liquids upto very high
pressures.
Comparing Equation (5.156) with Equation (5.157), we get
RT ln a = V (P – Po)
V
ln a = RT (P – Po) ... (5.158)

The concept of activity is especially useful in the study of solutions.


CHOICE OF STANDARD STATE
The standard state of a gas at any given temperature is that in which the fugacity of the gas is
unity, i.e., fo = 1.
f
Since a = fo , and fo = 1, ∴ a = f.

Therefore, activity of any gas becomes equal to its fugacity. For ideal gases, fugacity is equal
to the pressure.
The standard state for a liquid or a solid at any given temperature is the state of the pure
liquid or solid at a pressure of 1 atm.
EFFECT OF T AND P ON FUGACITY
We have, dG = RT d ln f
f
∆G = G – Go = RT ln fo ... (5.159)

where Go and fo refer to the Gibbs free energy and fugacity respectively at a very low
pressure where the gas approximates ideal behaviour.
Rearranging Equation (5.159) gives
f G Go
R ln fo = R ln f – R ln fo = T – T

Differentiating the above equation with respect to T at constant P gives


∂ ln f – R ∂ ln fo = ∂(G/T) – ∂(Go/T)
R   ∂T      ... (5.160)
 ∂T P  P  ∂T P  ∂T P
We know that the dependence of G on T is given by
∂(G/T) H
 ∂T  = – T2 ... Gibbs - Helmholtz equation
 P
and fo is equal to the gas pressure at very low pressures and hence is independent of temperature.
Chemical Engineering Thermodynamics - I 5.30 Thermodynamic Properties of Pure Fluids

With this, Equation (5.160) becomes

R
∂ ln f = – H – – Ho
 T2  T2 
 ∂T P
∂ ln f = Ho – H
i.e.,  ∂T  ... (5.161)
 P RT2
where H is the molar enthalpy of the gas at the given pressure P and Ho is the molar enthalpy at a
very low, i.e., zero pressure. Therefore, the difference Ho – H is the increase of molar enthalpy
accompanying the expansion of the gas from the pressure P to zero pressure at constant
temperature. Equation (5.161) expresses the dependence of fugacity on temperature at constant
pressure (i.e., effect of T on f at constant P).
The dependence of fugacity on pressure, at constant temperature may be expressed as
∂ ln f = V
 ∂P  ... (5.162)
 T RT
We know that dG = V dP = RT d ln f ... at constant T
Therefore, V dP = RT d ln f ... at constant T
V
d ln f = RT dP ... at constant T

∂ ln f V
∴  ∂P  = RT ... (5.163)
 T
Equation (5.163) indicates the effect of P on fugacity at constant T.
DETERMINATION OF FUGACITY OF REAL GASES
• RESIDUAL VOLUME METHOD FOR DETERMINING FUGACITY :
The change in the Gibbs free energy for any real gas at constant temperature is given by
dG = V dP = RT d ln f
∴ RT d ln f = V dP ... (5.164)
For a real gas, the quantity α (residual volume) which is a function of P and T is defined by
RT
α = V– P ... (5.165)

The residual volume (α) is the difference between the molar volume of the gas (V) and the
volume occupied by one mole of an ideal gas at the same T and P.
Rearranging Equation (5.165),
RT
V = α+ P ... (5.166)

Substituting Equation (5.166) in Equation (5.164), we get

RT d ln f = α + P  dP
RT
 
 α 1
d ln f = RT + P dP
 
Chemical Engineering Thermodynamics - I 5.31 Thermodynamic Properties of Pure Fluids

α 1 α
d ln f = RT dP + P dP = RT dP + d ln P

α
∴ d ln f – d ln P = RT dP

α
d ln (f/P) = RT dP ... (5.167)

Integrating this equation between a very low, virtually zero, pressure and a given pressure P
at constant temperature and remembering that f/P approaches unity as pressure approaches zero.
Therefore, ln f/P becomes zero at P = 0 (since ln 1 = 0).
f/P P

⌠ α
⌡ d ln f/P = ⌠
⌡ RT dP
1 0
P
α
ln f/P – ln 1 = ⌠
⌡ RT dP
0
P
α
ln f/P – 0 = ⌠
⌡ RT dP
0
P
1
i.e., ln f/P = RT ⌠
⌡ α dP ... (5.168)
0

To find the fugacity, we have to plot α (derived from experimentally determined molar
volumes of the gas) against pressure P. The area under the curve between P = 0 and P = P gives
the value of the integral on RHS of Equation (5.168).
Integral on the RHS = Area under the curve × (Scale of Y-axis) × (Scale of X-axis)
RT
We have, α = V– P

Dividing throughout by RT gives


α V 1
RT = RT – P
Taking 1/P common from the terms on the RHS,
α PV  1
RT = RT – 1 P
PV
This is useful when P v/s RT data are provided.

The value of the fugacity of the gas obtained may be less than or more than the pressure P.
(It is less than P at low pressures and more than P at very high pressures.)
P P
1 α
ln f/P = RT ⌠
⌡ α dP = ⌠
⌡ RT dP ... (5.169)
0 0
Chemical Engineering Thermodynamics - I 5.32 Thermodynamic Properties of Pure Fluids

Plot α/RT v/s P and find the area under the curve between P = 0 and P = P. The area gives
the value of the integral on RHS.
∴ ln f/P = Area under the curve × (Scale of X-axis) × (Scale of Y-axis)

Fig. 5.1 : Determination of fugacity using residual volume method


APPROXIMATE METHOD OF CALCULATION OF FUGACITY
At moderate pressure, the value of PV for any gas is a linear function of its pressure at
constant temperature. The functional relationship between PV and P may be written as
PV = RT + AP ... (5.170)
where A is a constant.
From this equation it is seen that α is given by
RT
α = V– P = A

so that α is constant over a range of moderate pressures.


P
1
We have, ln f/P = RT ⌠
⌡ α dP
0

αP
∴ ln f/P = RT ... (5.171)

At moderate pressures, f/P is close to unity and therefore, it is possible to make use of the
fact that ln x is approximately equal to x – 1 when x approaches unity; hence ln f/P = f/P – 1.
With this, Equation (5.171) becomes
f αP
P – 1 = RT

 αP  αP2
f = 1 + RT P = P + RT
 
Chemical Engineering Thermodynamics - I 5.33 Thermodynamic Properties of Pure Fluids

= P + V – P  RT
RT P2
 
VP2 RTP2 VP2
= P + RT2 – PRT = P + RT2 – P

VP2
f = RT2 ... (5.172)

This equation is used to calculate the fugacity of a gas from its pressure and molar volume at
moderate pressures.
GENERALISED METHOD FOR DETERMINING FUGACITY
The compressibility factor, Z, of a real gas is defined as
V Volume of the real gas
Z = RT/P = Volume of an ideal gas , at the same P and T

PV
Z = RT

V Z
∴ RT = P ... (5.173)

We have, RT d ln f = V dP, at constant T


V
d ln f = RT dP ... (5.174)

Substituting Equation (5.173) in Equation (5.174) gives


Z
d ln f = P dP
Adding and subtracting dP/P on the RHS of the above equation gives
Z dP dP Z dP dP
d ln f = P dP + P – P = P dP – P + P

dP dP
d ln f = (Z – 1) P + d ln P , since P = d ln P

dP
d ln f – d ln P = (Z – 1) P

(Z – 1)
d (ln f/P) = P dP

Integrating this equation between P = 0 and P = P, we get


[Remembering that f/P approaches unity as P approaches zero]
f/P P

⌠ (Z – 1)
⌡ d ln f/P = ⌠
⌡ P dP
1 0
P
(Z – 1)
ln f/P – ln 1 = ⌠
⌡ P dP
0
Chemical Engineering Thermodynamics - I 5.34 Thermodynamic Properties of Pure Fluids

P
(Z – 1)
ln f/P – 0 = ⌠
⌡ P dP
0
P
(Z – 1)
ln f/P = ⌠
⌡ P dP ... (5.175)
0

The value of the integral is found by plotting (Z – 1)/P against P.


The term dP/P is identical with dPr/Pr. Therefore, Equation (5.175) becomes
Pr
Z–1
ln f/P = ⌠
⌡ Pr dPr … (5.176)
0

Pr = P/Pc, where Pc is the critical pressure of the gas.


[Note that f/P approaches unity as Pr approaches zero.]
f/P Pr

⌠ (Z – 1)
⌡ d ln f/P = ⌠
⌡ Pr dPr
1 0
Pr
(Z – 1)
ln f/P – ln 1 = ⌠
⌡ Pr dPr
0
Pr
(Z – 1)
ln f/P = ⌠
⌡ Pr dPr ... (5.177)
0

DETERMINATION OF FUGACITY FROM EQUATION OF STATE


We know that dG = V dP = RT d ln f
Integrating the above equation between P = Po (very low pressure) at which f = fo and P = P
at which f = f, we get
P
f 1
ln fo = RT ⌠
⌡ V dP ... (5.178)
Po

If an equation of state which can be put in a form in which V can be expressed as function P
is available, then the integral in Equation (5.178) can be easily evaluated.
On the other hand, if P is expressed as a function of V, the variable in the integrand in
Equation (5.178) is changed by integrating by parts to evaluate the integral.

⌡ V dP = PV – ⌠
⌡ P dV
With this, the integral in Equation (5.178) becomes
P V


⌡ V dP = PV – PoVo – ⌠
⌡ P dV … (5.179)
Po Vo
Chemical Engineering Thermodynamics - I 5.35 Thermodynamic Properties of Pure Fluids

where Vo is the molar volume of the gas corresponding to the low pressure Po. Since at very low
pressures the gas behaves ideally, PoVo in the above equation can be replaced by
RT (as PoVo = RT). Therefore,
P V


⌡ V dP = PV – PoVo – ⌠
⌡ P dV ... (5.180)
Po Vo

1  
V

and ln f/fo = RT PV – RT ⌠


⌡ P dV  ... (5.181)
 
 Vo 
Remembering that f/P approaches unity at low pressure, i.e. fo/Po is virtually unity, it follows
that ln f/fo in Equation (5.181) may be replaced by ln f/Po.

1  
V

Therefore, ln f/Po = ln f – ln Po = RT PV – RT – ⌠


⌡ P dV ... (5.182)
 
 Vo 
If we express P as a function of V at constant T from a given equation of state, then the
integral in the above equation can be evaluated analytically.

FUGACITIES OF PURE SOLIDS AND LIQUIDS

Every liquid or solid has a definite vapour pressure at a given temperature, although in case
of non-volatile substances it may be immeasurably small. There is always a definite pressure at
which a liquid or solid is in equilibrium with its vapour, at constant temperature. When two
phases of a single substance are in equilibrium at a given T and P, the molar free energy in both
the phases is identical. [GL = GV]. Since G = RT ln f + C, the fugacity of the liquid (or solid)
will be the same as that of the vapour with which it is in equilibrium [fL = fV‚ fS = fV] . If the
pressure of the vapour is not too high, the fugacity of the vapour will be equal to the vapour
pressure. So the fugacity of a liquid (or solid) is approximately equal to its vapour pressure.

If the vapour pressure of a liquid (or solid) is very high at which the vapour cannot
approximate ideal behaviour, then its fugacity is obtained by means of
V (Ps)2
fs = RT ... (5.183)

where Ps is the saturation pressure and fs is the saturation fugacity or the fugacity at the
saturation pressure.

We have, V dP = RT d ln f
V
d ln f = RT dP
Chemical Engineering Thermodynamics - I 5.36 Thermodynamic Properties of Pure Fluids

Since liquids are incompressible fluid, the fugacity of a liquid at any pressure P is given by
f P

⌠ V
⌡ d ln f = RT ⌠
⌡ dP ... for incompressible fluid, V does change
fs Ps

f V
ln fs = RT (P – Ps) ... (5.184)

where V is the molar volume of the liquid. For the vapour behaving as an ideal gas,
fs = Ps = saturation pressure (vapour pressure).
DEPARTURE FUNCTIONS
The thermodynamic properties of a real fluid can be roughly estimated from a knowledge of
the departure functions or residual properties.
The residual property (or property departure) of a real fluid is defined as the difference
between the thermodynamic property of the fluid at the specified P and T and the property of the
same fluid in the ideal gas state (i.e., considering the fluid to be an ideal gas) at the same P & T.
If M is any extensive thermodynamic property, then the residual property M or the departure
function for M is defined as
MR = M – Mig ... (5.185)
where M is the thermodynamic property of the real gas at P and T and Mig is the thermodynamic
property of the real gas in the ideal gas state at the same P and T.
The residual enthalpy or enthalpy departure is defined as
HR = H – Hig ... (5.186)
where HR is the residual enthalpy.
The difference between the enthalpies of a real gas and the real gas in the ideal gas state at
the same P and T is called the enthalpy departure or residual enthalpy.
Similarly, the entropy departure or residual entropy is the difference between the entropies of
a real gas and the real gas in the ideal gas state at the same P and T.
SR = S – Sig ... (5.187)
S = Entropy of the real gas at the specified P and T
Sig = Entropy of the real gas in the ideal gas state at the same P and T
H = Enthalpy of the real gas at the specified P and T and
Hig = Enthalpy of the gas in the ideal gas state at the same P and T
Differentiating Equations (5.186) and (5.187) with respect to P at constant T yields
∂HR = ∂H – ∂Hig
 ∂P      ... (5.188)
 T  ∂P T  ∂P T
∂SR = ∂S – ∂Sig
 ∂P      ... (5.189)
 T ∂PT  ∂P T
Chemical Engineering Thermodynamics - I 5.37 Thermodynamic Properties of Pure Fluids

The variation of S with P at constant T is given by


∂S = – ∂V
∂P  ∂T  ... (5.190)
 T  P
The variation of H with P at constant T is given by
∂H = V – T ∂V
 ∂P   ∂T  ... (5.191)
 T  P
∂Sig and ∂Hig
Let us find    ∂P 
 ∂P T  T
For an ideal gas, PV = RT
RT
V = P
Differentiating w.r.t. T at constant P gives
∂V = R
 ∂T 
 P P
∂Sig ∂V
 ∂P  = –  ∂T  ... for an ideal gas from Equation (5.190)
 T  P
∂Sig = – R
∴  ∂P  … (5.192)
 T P
RT
V = P
Differentiating V w.r.t. P at constant T gives
∂V = R
 ∂T 
 P P
∂Hig = V – T ∂V … for an ideal gas
 ∂P   ∂P 
 T  T
∂Hig = V – T R = V – RT = V – V = 0
 ∂P  P
 T P
∂Hig
 ∂P  = 0 ... (5.193)
 T
So, for a real gas in the ideal gas state :
∂Hig = 0 and ∂Sig = – R
 ∂P   ∂P 
 T  T P
∂H ∂S
Let us find   and   for a real gas.
 ∂P T ∂PT
The compressibility factor is given by
V
Z = (RT/P) , where V is the molar volume

V = P = R  P  = R Z · P 
ZRT ZT T
   
Chemical Engineering Thermodynamics - I 5.38 Thermodynamic Properties of Pure Fluids

Differentiating with respect to T at constant P gives


∂V = R Z · 1 × 1 + T ∂Z 
 ∂T   P P ∂TP
 P 
∂V = R Z + T ∂Z 
 ∂T  P  ∂T  ... (5.194)
 P  P
Substituting Equation (5.194) in Equation (5.110), we get
∂S = – ∂V
∂P  ∂T  ... Equation (5.190)
 T  P
∂S = – R Z + T ∂Z 
∂P P  ∂T  ... (5.195)
 T  P
Substituting Equation (5.194) in Equation (5.191), we get

∂H = V – T ∂V
 ∂P   ∂T  ... Equation (5.191)
 T  P

∂H = V – RT Z + T ∂Z 
 ∂P  P  ∂T 
 T  P
RTZ RT2 ∂Z
= V– P – P  
∂TP
RT2 ∂Z
= V–V– P  
∂TP

∂H = – RT2 ∂Z


 ∂P  P ∂TP ... (5.196)
 T

Substituting Equations (5.196) and (5.190) in Equation (5.188), we get

∂HR RT2 ∂Z RT2 ∂Z


 ∂P  = – P ∂T – 0 = – P ∂T ... (5.197)
 T  P  P
Substituting Equations (5.192) and (5.195) in Equation (5.189), we get
∂SR = – R Z + T ∂Z  – – R
 ∂P  P  ∂T   P 
 T  P
∂SR R RT ∂Z R
 ∂P  = – P Z – P ∂T + P
 T  P

∂SR = – R (Z – 1) – RT ∂Z
∴  ∂P  P ∂TP ... (5.198)
 T P
Chemical Engineering Thermodynamics - I 5.39 Thermodynamic Properties of Pure Fluids

Integrating Equations (5.197) and (5.198) for an isothermal change from P = 0 (where HR = 0
and SR = 0) to P = P, we get
P

⌠ ∂Z dP
HR = – RT2 ⌡ ∂T · P ... (5.199)
0 P

P
 dP ∂Z dP
and SR = – R ⌠
⌡ (Z – 1) P + T ∂T · P  ... (5.200)
 0 P 
∂HR = –RT2 ∂Z
It is done as  ∂P  P ∂TP
 T
RT2 ∂Z
∴ dHR = – P   dP at constant T
∂TP
HR P

⌠ ∂Z · dP
⌡ dHR = – RT2 ⌠
⌡ ∂T P
 P
0 0
P
∂Z dP
HR – 0 = – RT2 ⌠
⌡ ∂T · P
P0
P
∂Z dP
∴ HR = – RT2 ⌠
⌡ ∂T · P
0

∂Z
The values of Z and   are calculated directly from experimental PVT data and the
∂TP
integrals in Equations (5.199) and (5.200) are evaluated by numerical or graphical methods.
Alternatively, when Z is expressed by an equation of state, the integrals of Equations (5.199) and
(5.200) are evaluated analytically. Therefore, whenever we are provided with PVT data or an
equation of state, we can evaluate HR and SR. Knowing HR and SR, we can obtain the other
departure functions or residual properties by using the relations
(U – Uig) = (H – Hig) – P (V – Vig) ⇒ UR = HR – PVR ... (5.201)
(G – Gig) = (H – Hig) – T (S – Sig) ⇒ GR = HR – TSR ... (5.202)
(A – Aig) = (U – Uig) – T (S – Sig) ⇒ AR = UR – TSR ... (5.203)
The residual volume or volume departure is
VR = V – Vig ... (5.204)

Therefore,
RT
VR = V – P , since Vig = RT
 P
V
Since Z = (RT/P)

ZRT
V = P ... for a real gas
Chemical Engineering Thermodynamics - I 5.40 Thermodynamic Properties of Pure Fluids

ZRT RT
Therefore, VR = P – P
RT
VR = P (Z – 1) ... (5.205)

Equation (5.205) relates the residual volume to the compressibility factor.


Knowing the residual enthalpy and entropy, one can calculate the enthalpy and entropy of the
real gas at the specified P and T using Equations (5.199) and (5.200). The enthalpy of the gas in
the ideal gas state at the same P and T for use in these two equations can be obtained by using the
following equations.
T

+⌠
ig
Hig = Ho ⌡ CP dT ... (5.206)
To
T
ig dT P
Sig = So + ⌠
⌡ CP T – R ln Po ... (5.207)
To

where Po and To are the reference state pressure and temperature, respectively where the enthalpy
and entropy values are Ho and So respectively.
THERMODYNAMIC DIAGRAMS
The thermodynamic properties of pure substances of significant concern are pressure (P),
temperature (T), volume (V), enthalpy (H) and entropy (S). The other properties such as U, A
and G can be calculated in terms of P, T, V, H and S.
The thermodynamic properties of a pure substance can be presented in the form of diagrams.
A thermodynamic diagram depicts the temperature, pressure, volume, enthalpy of a pure
substance on a single plot.
The thermodynamic diagrams in common use are :
(i) Pressure versus enthalpy (P-H diagram),
(ii) Temperature versus entropy (T-S diagram),
(iii) Enthalpy versus temperature (H-T diagram) and
(iv) Enthalpy versus entropy (H-S diagram / known as Mollier diagram).
These diagrams, providing properties in terms of two independent variables and on which
other properties are plotted as lines of constant property value like constant temperature lines,
etc., find wide applications in the calculation of thermodynamic properties and thermodynamic
analysis of processes (e.g., use of T-S diagrams in the analysis of refrigeration cycles).
1. P-H diagram : A typical (pressure - enthalpy) P-H diagram is represented in Fig. 5.2,
wherein ln P is plotted as an ordinate and H as an abscissa and a family of constant temperature
lines, constant volume lines and constant entropy lines are plotted.
Chemical Engineering Thermodynamics - I 5.41 Thermodynamic Properties of Pure Fluids

Constant S

In P Constant V

Liquid
Liquid + Vapour Vapour

Constant T

Enthalpy, H
Fig. 5.2 : Pressure - enthalpy (P-H) diagram
This diagram is especially used for calculating the heat loads and temperature changes in
refrigeration systems. Since the refrigeration cycles have both constant pressure (evaporators and
condensers) and constant enthalpy (throttling valve) processes, P - H diagrams are useful in
analysing them.
The area under the dome-shaped envelope is the two-phase region. Temperature lines are
almost vertical in the liquid-phase region and horizontal in the two-phase region. These lines
drop steeply in the vapour-phase region. Horizontal lines in the two-phase region represent
constant pressure and constant temperature lines. The constant entropy lines, shown as
continuous lines in the vapour phase and two-phase regions are tilted from the Y-axis.
2. H-T diagram : A typical H-T diagram is represented in Fig. 5.3 wherein H

Saturated
Constant S
vapour

Constant P
Enthalpy, H

Constant Critical point


quality

Saturated liquid

Temperature, T
Fig. 5.3 : H-T diagram
Chemical Engineering Thermodynamics - I 5.42 Thermodynamic Properties of Pure Fluids

(enthalpy) is plotted as an ordinate and T (temperature) is plotted as an abscissa. It consists of a


family of constant pressure lines, constant quality lines and constant entropy lines plotted on
enthalpy - temperature coordinates.
These diagrams are useful in analysing the throttling as well as constant pressure flow
processes. The vertical distance between the saturated vapour curve and the saturated liquid
curve gives the value of the latent heat of vaporisation at a particular temperature and pressure.
3. T-S diagram : A typical T-S diagram is represented in Fig. 5.4, wherein T is plotted as
an ordinate and S is plotted as an abscissa. It consists of a family of constant enthalpy lines,
constant volume lines, constant quality lines and constant pressure lines plotted on temperature -
entropy coordinates. The area under the dome-shaped envelope is the two-phase region.

Fig. 5.4 : T-S diagram


A reversible adiabatic process is called an isentropic process (during this process S is
constant). Therefore, it can be represented by a vertical line on a T-S diagram. Hence these
diagrams are useful to follow the temperature changes during isentropic processes. In the case of
a reversible adiabatic expansion process (turbines) or a reversible adiabatic compression process
(compressors), a vertical line drawn from the initial pressure to the final pressure on this diagram
shows the path followed by the fluid. The horizontal distance between the saturated vapour curve
and the saturated liquid curve gives the value of λ/T, where λ is the latent heat of vaporization
at T.
4. H-S diagram (Mollier diagram) : A sketch of the H-S diagram is shown in Fig. 5.5.
It consists of a family of constant temperature lines, constant quality lines and constant pressure
lines plotted on enthalpy - entropy coordinates. This diagram is known as the Mollier diagram.
Chemical Engineering Thermodynamics - I 5.43 Thermodynamic Properties of Pure Fluids

Constant superheat
Constant P

Saturated Constant T
vapour Vapour

Critical
Enthalpy, H

point Constant
T and P

Liquid ur
po Constant
Va quantity
+
uid
Saturated Liq
liquid Triple-point line

Entropy, S

Fig. 5.5 : H-S diagram


The constant temperature lines and constant pressure lines shown within the two-phase
region, separate in the vapour-phase region into pressure lines which rise continuously and
temperature lines which drop, ultimately become horizontal. The lines for constant temperature
and constant superheat appear in the vapour-phase region. The term superheat denotes the
difference between the actual temperature and the saturation temperature at the same P. These
diagrams are useful to determine the energy requirements in flow processes as well as the
temperature changes in isenthalpic and isentropic processes.

SOLVED EXAMPLES
Example 5.1 :

Show that T 
∂P – P = ∂U
  ∂V 
∂TV  T
Solution :
For S = f (V, T), the total differential is
 ∂S  ∂S
dS =   dV +   dT ... (1)
 
∂V ∂T
T V
Chemical Engineering Thermodynamics - I 5.44 Thermodynamic Properties of Pure Fluids

 ∂S  = ∂P
We have, ∂V   ... Maxwell’s equation
 T ∂TV
 ∂S 
Substituting for   , Equation (1) becomes
∂V T

∂P ∂S
dS =   dV +   dT ... (2)
 
∂T
V
∂T V

We have, dU = T dS – P dV ... (3)


Substituting for dS from Equation (2) into Equation (3) yields
∂P ∂S 
dU = T   dV +   dT – P dV ... (4)
∂T V
∂T V 
∂S = CV
We have, ∂T ... (5)
 V T

Equation (4) becomes


∂P CV
dU = T   dV + T T · dT – P dV
∂T V

∂P
∴ dU = T   dV + CV dT – P dV
∂T V

Differentiating the above equation with V at constant T gives


∂U = T ∂P + 0 – P
∂V ∂T
 T  V
∂U = T ∂P – P
∴ ∂V ∂T
 T  V

T   – P =  
∂P ∂U
... Ans.
∂TV ∂VT
Example 5.2 :
 (CP – CV) κ
Show that dH = CP dT + V –  dP
 β 
Solution :
We have, dH = T dS + V dP ... (1)
For S = f (P, T), the total differential of S is
∂S ∂S
dS =   dP +   dT ... (2)
∂PT
∂T P
Chemical Engineering Thermodynamics - I 5.45 Thermodynamic Properties of Pure Fluids

∂Q
We have, CP =   and dQ = T dS
 ∂T  P


∂S
CP = T  
∂T P

Rearranging, we get

∂S = CP
∂T ... (3)
 P T

Substituting Equation (3) into Equation (2) gives

∂S CP
dS =   dP + T dT ... (4)
 
∂P
T

∂S ∂V
We have, –  =   ... Maxwell’s equation
∂PT  ∂T P

∂S
Substituting for   , Equation (4) becomes
∂P T

∂V CP
dS = –   dP + T dT
 ∂T  P

CP ∂V
dS = T dT –   dP ... (5)
 ∂T  P

Substituting Equation (5) into Equation (1), we get

CP ∂V 
dH = T  T dT –   dP + V dP
  ∂T  P 
 ∂V 
dH = CP dT + V – T    dP ... (6)
  ∂T   P

1 ∂V
We have, β = V  ... Coefficient of expansion
 ∂T P

1 ∂V
κ = –V  ... Coefficient of compressibility
 ∂P T
2
∂V  ∂P 
CP – CV = – T   ∂V
 ∂T P  T
Chemical Engineering Thermodynamics - I 5.46 Thermodynamic Properties of Pure Fluids

2
 ∂V  ∂P    1 ∂V 
– T  ∂T P ∂V  – V  ∂P  
(CP – CV) κ  T  T
∴ =
β 1 ∂V
V  ∂T 
P
2
∂V  ∂P 
T   
 ∂T P ∂V T ∂V 1
= ... as   =
∂V  ∂P   T  
∂P ∂P
 ∂T  ∂V ∂V
   P T
  T

(CP – CV) κ ∂V


= T  ... (7)
β  ∂T P
∂V
Substituting for T   from Equation (7) into Equation (6), we get
 ∂T  P

 (CP – CV) κ
dH = CP dT + V –  dP ... Ans.
 β 
Example 5.3 :
∂U = – T ∂V + P ∂V 
Show that  ∂P    ∂T   ∂P  
 T  P
 T
Solution :
We have, dU = T dS – P dV ... (1)
Let S = f (V, T). The total differential of S is
∂S  ∂S 
dS =   dT +   dV … (2)
∂T V
∂V T

 ∂S  ∂P
We have, ∂V = ∂T ... Maxwell’s relation
 T  V
∂S ∂S Cv
and CV = T   ∴   = T
∂T V
∂TV
With these values of partial derivative terms, Equation (2) becomes
CV
dS = T dT +   dV
∂P ... (3)
∂T V
Substituting for dS from Equation (3) into Equation (1), we get
CV ∂P 
dU = T  T dT +   dV – P dV
 ∂T   V

∂P
dU = CV dT + T   dV – P dV
∂T V
Chemical Engineering Thermodynamics - I 5.47 Thermodynamic Properties of Pure Fluids

∂V
 ∂T 
∂P = – ∂V  ∂P  = –  P
We have, ∂T  ∂T  ∂V
 V  P  T ∂V
 ∂P 
 T
∂V
T 
 ∂T  P
∴ dU = CV dT – · dV – P dV
∂V
 ∂P 
  T
Dividing by dP and imposing the constant temperature restriction, the above equation
becomes
∂V
 ∂T 
∂U = 0 – T  P · ∂V – P ∂V
 ∂P     ∂P 
 T ∂V  ∂P T  T
 ∂P 
  T

∂U = – T ∂V – P ∂V 


 ∂P   ∂T  ∂P   ... Ans.
 T  P
 T
Example 5.4 :
∂V
∂T
 P
Show that CP = T
∂T
∂P
 S
Solution :
∂Q
We have, CP =  
 ∂T 
P
Since dQ = T dS
∂S
CP = T   ... (1)
∂T P
Let S = f (T, P)
The total differentiation of S is
∂S ∂S
dS =   dT +   dP ... (2)
∂T P
∂P T

At constant S, dS = 0, so Equation (2) becomes


∂S ∂S
0 =   dT +   dP
 
∂T
P
∂P T

∂S dP = – ∂S dT
∂P ∂T
 T  P
Chemical Engineering Thermodynamics - I 5.48 Thermodynamic Properties of Pure Fluids

Dividing by dT and imposing the restriction of constant S, the above equation becomes
∂S ∂P = – ∂S
∂P ∂T ∂T
 T  S  P
On rearrangement it gives
∂S
∂T
∂P = –  P
∂T
 S ∂S
∂P
 T

∂S = – ∂P ∂S


∴ ∂T ∂T ∂P ... (3)
 P  S  T
∂S = – ∂V
We have, ∂P  ∂T  ... Maxwell’s equation
 T  P
Substituting this in Equation (3) gives
∂S = – ∂P – ∂V 
∂T ∂T   ∂T  
  P
    
S P

∂S = ∂P ∂V


∂T    
 P ∂TS  ∂T P
∂S
Substituting for   , Equation (1) becomes
∂T P

∂P ∂V
CP = T     ... (4)
∂T  ∂T 
S P

∂P = 1
We have, ∂T
 S ∂T
∂P
 S
Substituting this in Equation (4), we get
∂V
T 
∂TP
CP = ... Ans.
∂T
 ∂P 
  S

Example 5.5 :
Using the method of partial derivatives, express the following partial derivatives in terms of
measurable quantities :
∂H ∂H ∂H
(i)   (ii)   and (iii)  
 ∂P  T
∂V P
 ∂T  V
Chemical Engineering Thermodynamics - I 5.49 Thermodynamic Properties of Pure Fluids

Solution :
∂H
(i)  
 ∂P  T
The differential equation for H is
dH = T dS + V dP
Dividing the above equation by dP holding T constant, we get
∂H = T ∂S + V
 ∂P  ∂P ... (1)
  T
  T

Substituting the Maxwell’s relation given by


∂S ∂V
–  =  
∂PT  ∂T P
in Equation (1), we get
∂H = – T ∂V + V
 ∂P   ∂T  ... (2)
 T  P
The partial derivative on the R.H.S. is related to β. The coefficient of volume expansion is
given by
1 ∂V
β = V 
 ∂T  P

∂V = Vβ
∴  ∂T  ... (3)
 P
∂V
Substituting for   in Equation (2) gives
 ∂T  P

∂H = – βVT + V
 ∂P 
 T
∂H = – V (β
 ∂P  β T – 1) = V (1 – β T) ... Ans. (i)
 T

(ii)  
∂H
 ∂V P
The differential equation for enthalpy is
dH = T dS + V dP ... (1)
Consider S = f (V, T)
The total differential of S is
 ∂S  ∂S
dS =   dV +   dT ... (2)
∂V T
∂T V
The first partial derivative is related to the Maxwell's relation and the second one to the
definition of CV.
Chemical Engineering Thermodynamics - I 5.50 Thermodynamic Properties of Pure Fluids

∂S
We know that, CV = T  
∂T V

∂S = CV
∂T
 V T

 ∂S  = ∂P
∂V   ... Maxwell's relation
 T ∂TV
Substituting for the partial derivatives, Equation (2) becomes
∂P CV
dS =   dV + T dT ... (3)
 
∂T
V

The cyclic relation among P, T and V is


∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P 
     
V P T

∂V
 ∂T 
∂P = – 1  P 1 ∂y
∂T = – ... as =  
 V   ∂V
∂T ∂V  
∂x ∂xz
∂V  ∂P   ∂P  ∂y
 P  T  T  z
∂P
Substituting for   in Equation (3) gives
∂T V

∂V
 ∂T 
 P CV
dS = – dV + T dT ... (4)
∂V
 ∂P 
  T

1 ∂V ∂V
We know that, β = V  ∴   = βV
 ∂T P  ∂T P
1 ∂V ∂V
κ = –V  ∴   = – κV
 ∂P  T
 ∂P  T

Substituting for the partial derivatives, Equation (4) becomes


– βV CV
dS = dV + T dT
– κV
βT
∴ T dS = dV + CV dT ... (5)
κ
Substituting Equation (5) in Equation (1), we get
βT
dH = dV + CV dT + V dP
κ
Chemical Engineering Thermodynamics - I 5.51 Thermodynamic Properties of Pure Fluids

Dividing by dV holding P constant, we obtain


∂H = βT + C  ∂T 
∂V V   +0
  κ
P
∂V P

∂H = βT + CV ∂x 1
∂V ... since   = ... (6)
 P κ ∂V ∂yz ∂y
 ∂T  ∂x
 P   z
1 ∂V
We know that, β = V 
 ∂T  P

∂V = βV
 ∂T 
  P

∂V
Substituting for   in Equation (6), we get
 ∂T  P

∂H = β T + CV
 ∂V  ... (7) ... Ans. (ii)
 P κ βV
∂H can be expressed in terms of C instead of C . (This will reduce the measurable
∂V
 P P V

quantities.) Let us do this.


β 2 VT
We know that, CP – CV =
κ
β 2 VT
CV = CP –
κ
Substituting for CV in Equation (7), we get
∂H = βT + CP – β 2 VT/κ
∂V
  P
κ βV
∂H = βT + CP – β 2 VT
∂V
 P κ βV κβV

∂H = βT + CP – βT
∂V
 P κ βV κ

∂H = CP
∴  ∂V  ... Ans. (ii)
 P βV

(iii)   : We know that,T dS =


∂H βT
dV + CV dT ... (5)
 ∂T 
V
κ
We know that, dH = T dS + V dP
Substituting for T dS in the above equation gives
βT
dH = dV + CV dT + V dP ... (1)
κ
Chemical Engineering Thermodynamics - I 5.52 Thermodynamic Properties of Pure Fluids

Dividing Equation (1) by dT and imposing the constant volume (V) restriction, Equation (1)
becomes
∂H = 0 + C + V ∂P
 ∂T  ∂T
 V V
 V

∂H = C + V ∂P
 ∂T  ∂T ... (2)
 V V
 V
We know that,
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Cyclic relation
 V  P  T

∂V
 ∂T 
∂P = – 1  P
∂T = – ... Using reciprocal relation
 V  ∂T  ∂V ∂V
∂V  ∂P   ∂P 
 P  T  T

∂P
Substituting for   in Equation (2) gives
∂T V

∂V
 ∂T 
∂H = C – V  P
 ∂T  ... (3)
 V V
∂V
 ∂P 
 T
1 ∂V
We know that, β = V  ... Coefficient of volume expansion
 ∂T 
P

∂V = βV
∴  ∂T 
 P
1 ∂V
κ = –V  ... Coefficient of compressibility
 ∂P T
∂V = – κV
∴  ∂P 
 T
Substituting for the partial derivative terms of Equation (2), we get
∂H = C – V βV
 ∂T 
 V V
(– κV)

∂H = C + β V
 ∂T  ... Ans. (iii)
 V V
κ
Chemical Engineering Thermodynamics - I 5.53 Thermodynamic Properties of Pure Fluids

Example 5.6 :
Prove the following :
∂CV ∂2P
  = T ∂T2
 ∂V   VT

Solution :
The heat capacity at constant volume is given by

CV = T  
∂S
∂T V

∂S = CV
∂T ... (1)
  V
T
Differentiate Equation (1) with respect to V keeping T constant. This gives
∂2S ∂CV/T 1 ∂CV
=   = T  ... (2)
∂V ∂T  ∂V   ∂V 
T T

∂P =  ∂S 
We have, ∂T ∂V ... Maxwell’s relation
 V  T
Differentiate the above equation with respect to T keeping V constant. This gives
∂2P = ∂2S
∂T2 ... (3)
  V
∂T ∂V
∂2z ∂2z
We know that =
∂x ∂y ∂y ∂x
∂S
2 ∂2S
∴ =
∂T ∂V ∂V ∂T
Substituting for the mixed partial derivatives from Equations (3) and (1), the above relation
yields
∂2P = 1 ∂CV
∂T2
  T  ∂V 
V T

∂CV
 = T  2
∂ 2P
∴  ... Ans.
 ∂V T ∂T V
Example 5.7 :
∂CP ∂2V
Prove the following : (i)   = – T  2
 ∂P T ∂T P
Solution :
The heat capacity at constant pressure is given by
∂S
CP = T  
∂T P

∂S = CP
∂T ... (1)
 P T
Chemical Engineering Thermodynamics - I 5.54 Thermodynamic Properties of Pure Fluids

Differentiate Equation (1) with respect to pressure keeping T constant. This gives
∂2S ∂CP/T 1 ∂CP
=   = T  ∂P  ... (2)
∂P ∂T  ∂P T  T

∂S ∂V
–  =  
We have, ... Maxwell’s relation
∂PT  ∂T P

∂S = – ∂V
∂P  ∂T  ... (3)
 T  P
Differentiate Equation (3) with respect to temperature (T) keeping P constant. This gives
∂2S ∂2V
= –  2 ... (4)
∂T ∂P  ∂T P
∂2z ∂2z
We know that =
∂x ∂y ∂y ∂x
∂2S ∂2S
∴ =
∂P ∂T ∂T ∂P
Substituting for the mixed partial derivatives from Equations (2) and (4), the above equation
yields
1 ∂CP ∂2V
 
T  ∂P  = –  ∂T2 
T
 P

∂CP
  = – T  2 
∂ 2V
∴ ... Ans.
 ∂P T ∂T P
Example 5.8 :
Derive the following relation using the partial derivatives method :
κ CP
T dS = CV dP + dV
β Vβ
Solution :
κ CP
T dS = C dP + dV
β V Vβ
We require the relation in terms of the independent variables P and V. Therefore, consider
S = f (P , V).
The total differential of S is
∂S  ∂S 
dS =   dP +   dV ... (1)
∂P V
∂V P
Chemical Engineering Thermodynamics - I 5.55 Thermodynamic Properties of Pure Fluids

∂S
Consider  
∂P V

∂S = 1
∂P
 V ∂P
∂S
 V
Substituting the Maxwell’s relation given by
∂P  ∂T 
–  =  
∂S ∂V
V S
in the above equation, we get
∂S = – 1
∂P ... (2)
 V  ∂T 
∂V
  S
We know that,
 ∂T  ∂V ∂S = – 1
∂V  ∂S  ∂T ... Cyclic relation among T, V and S
     
S T V

 ∂T  –1
∴ ∂V = ∂V ∂S
 S    
 ∂S  ∂T
 T  V
 ∂S 
∂V
 ∂T  = –  T 1  ∂S 
∂V , ... as =   ... Reciprocal relation
 S ∂S ∂V ∂VT
∂T  ∂S 
 V   T

 ∂T 
Substituting for   , Equation (2) gives
∂V S

∂S
∂T
∂S  V
∂P = ∂S
 V  
∂V
 T
Substituting the Maxwell's relation given by
 ∂S  ∂P
∂V = ∂T
   
T V
in the above equation, we get
∂S
∂T
∂S =  V
∂P ... (3)
 V ∂P
∂T
 V
Chemical Engineering Thermodynamics - I 5.56 Thermodynamic Properties of Pure Fluids

We know that,
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Cyclic relation among P, T and V
 V  P  T
∂P = –1
∂T
 V   ∂V
∂T
∂V  ∂P 
 P  T
∂V
 ∂T 
 P 1 ∂V
= – as =  
∂V  ∂T   ∂T P
 ∂P  ∂V
 T   P
... Reciprocal relation between T and V
∂P
Substituting this value of   in Equation (3), we get
∂T V

∂S
∂T
∂S = –  V
∂P
 V ∂V / ∂V
 ∂T   ∂P 
 P  T
∂V ∂S
 ∂P  ∂T
∂S = –  T  V
∂P ... (4)
 V ∂V
 ∂T 
 P
1 ∂V 1 ∂V
We have, β = V  , κ = –V 
 ∂T P  ∂P T
∂S
and CV = T  
∂T V
1 ∂V ∂S 
–V   
κ CV  ∂P  ∂T
T V
∴ =
β T 1 ∂V
V  ∂T 
P

∂V ∂S
 ∂P  ∂T
κCV  T  V
= – ... (5)
βT ∂V
 ∂T 
 P
Combining Equations (4) and (5), we get
∂S = κCV
∂P ... (6)
  βT
V
Chemical Engineering Thermodynamics - I 5.57 Thermodynamic Properties of Pure Fluids

Substituting Equation (6) in Equation (1), we get


CV  ∂S 
dS = κ dP +   dV ... (7)
βT ∂V P

 ∂S 
Let us express   in terms of CP, β and V.
∂V P

 ∂S   ∂S  = 1
Consider   . ∂V ... Reciprocal relation ... (8)
∂V  P ∂V
P  ∂S 
 P
Substituting the Maxwell's relation given by
∂V = ∂T
 ∂S  ∂P
   
P S
in Equation (8), we get
 ∂S  = 1
∂V ... (9)
 P ∂T
∂P
  S
We know that,
∂T ∂P ∂S = – 1
∂P ∂S ∂T ... Cyclic relation among T, P and S
     
S T P

∂T = – 1
∂P
 S   ∂P
∂S
∂T ∂S
 P  T
∂T
Substituting for   in Equation (9), we get
∂P S

 ∂S  = – ∂S ∂P
∂V ∂T ∂S
 P  P  T
∂S
∂T
 P ∂P
= – ... Using reciprocal relation for   ... (10)
∂S ∂S
∂P T
  T

∂S ∂V
We know that, –  =   ... Maxwell's relation
∂PT  ∂T P
Substituting this Maxwell's relation, Equation (10) becomes
∂S
– 
∂S
∂T
 ∂S  = ∂T P
 P
∂V = ... (11)
 P ∂V ∂V
–   ∂T 
 
∂T   P P
Chemical Engineering Thermodynamics - I 5.58 Thermodynamic Properties of Pure Fluids

The partial derivative in the numerator is related to CP and that in the denominator is related
to β.
We know that,
∂S
CP = T  
∂T P

∂S = CP
∴ ∂T ... (12)
 P T

1 ∂V
β = V 
 ∂T P
∂V = Vβ
∴  ∂T  ... (13)
 P
Combining Equations (11), (12) and (13), we get
 ∂S  = CP  1  = CP
∂V T Vβ  ... (14)
 P TVβ

Substituting Equation (14) into Equation (7) gives


κCV CP
dS = dP + dV
βT TVβ
κCV CP
∴ T dS = dP + dV ... Ans.
β β

Example 5.9 :
Show that for a gas obeying van der Waals equation of state,
R
CP – CV =
1 – [2a (V – b)2 / (RTV3)]
where a and b are van der Waals constants.
Solution :
The van der Waals equation of state is
P + a  (V – b) = RT
 V 2
Differentiating the above equation with respect to V at constant T gives

P + a  (1) + (V – b)  ∂P  + a (– 2V–3) = 0


 V 2 ∂V 
 T 
P + a  + (V – b)  ∂P  – 2a  = 0
 V 2 ∂V V3
 T 
Chemical Engineering Thermodynamics - I 5.59 Thermodynamic Properties of Pure Fluids

Rearranging gives
 ∂P  2a (V – b) 
– P + V 2
a
(V – b)   =
∂V V3  
T

P + a 
 ∂P  = 2a –  V 2
∂V
 T V3 (V – b)
2a (P + a/V2)
= V3 – (V – b)

Multiply and divide the last term on the R.H.S. by (V – b) to get


 ∂P  2a (P + a/V2) (V – b)
∂V = V3 –
 T (V – b)2

We have, P + V2 (V – b) = RT. With this and combining the terms, the above equation
a
 
becomes
 ∂P  = 2a (V – b)2 – RTV3
∂V ... (1)
 T V3 (V – b)2

Differentiate the van der Waals equation with respect to T at constant P :


P + a  ∂V + (V – b) 0 – 2a ∂V  = R
 V2  ∂T   V3  ∂T  
P   P

∂V P + a  – 2a (V – b) = R


 ∂T   V2 
 P V3

∂V = R
 ∂T 
 P P +  – 2a (V – b)
a
 V 2 V3
Multiply and divide the first term in the numerator of the above equation by (V – b).
∂V R
 ∂T  =
 P P + a  (V – b)
 V 2 2a (V – b)
(V – b) – V3

= RT
R  a
2a (V – b) as P + V2 (V – b) = RT
(V – b) – V3
R (V – b) V3
= RTV3 – 2a (V – b)2

∂V = – R (V – b) V3
 ∂T  ... (2)
 P 2a (V – b)2 – RTV3
Chemical Engineering Thermodynamics - I 5.60 Thermodynamic Properties of Pure Fluids

∂S ∂S
We know that CP = T   and CV = T  
∂T P
∂T
V

Consider S = f (P, T)
The total differential of S is
∂S ∂S
dS =   dP +   dT
∂P T
∂T P

∂S ∂S
Substituting for   from the definition of CP and for   from the Maxwell's relation
∂T P
∂P T

∂S = ∂V
∂P  
 T  ∂T P
in Equation (3), we get
∂V CP
dS = –   dP + T dT
 
∂T
P

Dividing by dT holding V constant, the above equation becomes


∂S ∂V ∂P CP
∂T = –  ∂T  ∂T + T
     
V P V

∂S
Substituting for   from the definition of CV gives
∂T V

CV ∂V ∂P + CP
= –  ∂T  ∂T
T  P  V T
Rearranging, we get
∂V ∂P 
CP – CV = T      ... (3)
 ∂T  ∂T 
P V

We know that,
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Reciprocal relation
 V  P  T
∂P = –1 ∂V  ∂P 
∂T = –   
 V  ∂T  ∂V  ∂T P ∂VT
∂V  ∂P 
 P  T
∂P
Substituting for   in Equation (3), we get
∂T V
2
∂V  ∂P  
CP – CV = – T      ... (4)
 ∂T P ∂V  T
Chemical Engineering Thermodynamics - I 5.61 Thermodynamic Properties of Pure Fluids

Substituting Equations (1) and (2) into Equation (4), we get


 –R (V – b) V3 2 2a (V – b)2 – RTV3 
CP – CV = – T  2a (V – b)2 – RTV3  
  V3 (V – b)2   

= – T 2a (V – b)2 – RTV3


R2V3
 
Taking RTV3 common from the denominator, we get
– TR2V3
CP – CV =
RTV3  RTV3 – 1
2a (V – b)2
 
– TR2V3
=
– RTV3 1 – RTV3 
2a (V – b)2
 
R
CP – CV = 2a (V – b)2 ... Ans.
1 – RTV3

R
CP – CV = ... Ans.
1 –  RTV3 
2a (V – b)2
 
Example 5.10 :
Show that for ideal gases, CP – CV = R
Solution :
We know that,
∂S ∂S
CP = T   and CV = T  
∂T P
∂T V

Consider S = f (P, T)
The total differential of S is
∂S ∂S
dS =   dP +   dT
∂P ∂T
T P

∂S ∂S
Substituting for   from the definition of CP and for   from the Maxwell's relation
 
∂T
P
∂P T

∂S = – ∂V in the above equation, we get


∂P  ∂T 
 T  P

∂V CP
dS = –   dP + T dT
 ∂T  P

CP ∂V
dS = T dT –   dP
 ∂T  P
Chemical Engineering Thermodynamics - I 5.62 Thermodynamic Properties of Pure Fluids

Dividing by dT holding V constant, we get


∂S = CP – ∂V ∂P
∂T T  ∂T  ∂T
 V P V

∂S
Substituting for   from the definition of Cv, the above equation becomes
∂T V

CV CP ∂V ∂P 
T = T –  ∂T  ∂T
P V

CP CV ∂V ∂P
T – T =  ∂T  ∂T
P V

∂V ∂P 
CP – CV = T      ... (1)
 ∂T  ∂T 
P V

For ideal gases : PV = RT


RT ∂V = R
V = P ∴  ∂T 
 P P

RT ∂P = R
P = V ∴ ∂T
 V V

Substituting for the partial derivatives in Equation (1), we get

CP – CV = T  P  V
R R
  
R2T R2
CP – CV = PV = (PV/T)

PV
We have, T = R With this, the above equation becomes
Cp – CV = R2/R = R
∴ CP – CV = R ... Ans.
Example 5.11 :
Develop an equation for evaluating the change in internal energy for process involving ideal
gases.
Solution :
Let the internal energy U be expressed as a function of T and V.
U = f (T, V)
The total differential of U is given as
∂U ∂U
dU =   dT +   dV ... (1)
 
∂T
V
∂V T
Chemical Engineering Thermodynamics - I 5.63 Thermodynamic Properties of Pure Fluids

The heat capacity at constant volume is given by


∂U
CV =  
 ∂T  V

∂U
Substituting for   in Equation (1), we get
 ∂T  V

∂U
dU = CV dT +   dV ... (2)
∂V T

Now we have to find the second term on the R.H.S. of Equation (2) for an ideal gas.
The differential equation for internal energy (U) is
dU = T dS – P dV
Dividing the above equation by dV holding T constant, we get
∂U = T  ∂S  – P
∂V ∂V ... (3)
  T
  T

Substituting the Maxwell's relation given by


 ∂S  = ∂P
∂V ∂T
   
T V
in Equation (3), we get
∂U = T ∂P – P
∂V ∂T ... (4)
 T  V
We know that
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Cyclic relation among P, T and V
     
V P T

∂P = –1
∴ ∂T
 V  ∂T  ∂V
∂V  ∂P 
 P  T
∂V
 ∂T 
 P ∂V 1
= – as   =
∂V  ∂T P  ∂T 
 ∂P  ∂V
 T   P

∂P
Substituting this value of   in Equation (4), we get
∂T V

∂V
–T 
∂U =  ∂T  P
∂V –P ... (5)
 T ∂V
 ∂P 
  T
Chemical Engineering Thermodynamics - I 5.64 Thermodynamic Properties of Pure Fluids

∂U
Substituting for   from Equation (5) in Equation (2) gives
∂V T

T ∂V 
∂T 

dU = CV dT –
 P
+P
 dV ... (6)
 ∂V 
 
  ∂P
T 
For 1 mole of an ideal gas :
PV = RT

V = P T
R
 
∂V = R , differentiating w.r.t. T at constant P
∴  ∂T 
 P P

PV = RT
RT
∴ V = P

∂V = –RT , differentiating w.r.t. P at constant T


∴  ∂P 
 T P2

Substituting for the partial derivatives of Equation (6) by their values obtained above, we get
dU = CV dT – – RT/P2 + P dV
T (R/P)
 
= CV dT –  RTP + P dV
2
–RT (P )
 
= CV dT – [– P + P] dV
∴ dU = CV dT ... only for an ideal gas ... Ans.
i.e., U = f (T) ... only for an ideal gas
Extra : We have,

T ∂V 
∂T 

dU = CV dT –  P
+ P dV ... (6)
 ∂V 
 
  ∂P  T 
1 ∂V
We know that β = V   ... Coefficient of volume expansion
 ∂T P
1 ∂V
κ = –V  ... Coefficient of compressibility
 ∂P  T
Chemical Engineering Thermodynamics - I 5.65 Thermodynamic Properties of Pure Fluids

∂V = βV
∴  ∂T 
 P
∂V = – κV
and  ∂P 
 T
Substituting for the partial derivatives, Equation (6) becomes
TβV + P dV
dU = CV dT –  
– κV 
βT
dU = CV dT + dV – P dV
κ
βT 
dU = CV dT +  – P dV ... for any gas ... (7)
κ 
RT ∂V = R
For an ideal gas, PV = RT ... V = P ...  ∂T 
 P P
1 ∂V R 1
∴ β = V   = VP = T
 ∂T  P

1 ∂V
κ = –V 
 ∂P T
RT ∂V = – RT
PV = RT ∴ V = P ∴  ∂T 
 T P2

κ = – V – P2  =  V  P2 = P P2 = P
1 RT RT 1 1 1

     
Substituting for β and κ in Equation (7), we get

dU = CV dT + T (1/P) – P dV
1 (T)
  
dU = CV dT + [P – P] dV
∴ dU = CV dT [... U = f (T) only for an ideal gas] ... Ans.
Example 5.12 :
Derive the following relation using the partial derivative method :
V
µ = µJT = C (Tβ – 1)
P

Solution :
The Joule-Thomson coefficient is defined as
∂T
µ =  
∂P H

We know, dH = T dS + V dP ... (1)


Chemical Engineering Thermodynamics - I 5.66 Thermodynamic Properties of Pure Fluids

Dividing by dP and holding T constant, the above equation becomes


∂H = T ∂S + V
 ∂P  ∂P ... (2)
  T
  T
Substituting the Maxwell's relation given by
∂S ∂V
–  =  
∂P  ∂T 
T P
in Equation (2), we get
∂H = – T ∂V + V
 ∂P   ∂T  … (3)
  T
  P
The cyclic relation for f (H, T, P) = 0 is
∂H ∂T  ∂P  = – 1
 ∂T  ∂P ∂H
     
P H T


∂T
µ =   = –
1
∂PH ∂H  ∂P 
 ∂T  ∂H
    P T

∂H
 ∂P 
  T  ∂P  1
= – , since   = … (4)
∂H  T ∂H
∂H
 ∂T   ∂P 
 P   T

∂H = C
We have,  ∂T  … (5)
 P P

Substituting for the partial derivative terms of Equation (4) from Equations (3) and (5),
we get
∂V 
 ∂T  + V
 P  1 ∂V ∂V
µ = – , We have : β = V   ∴ Vβ =  
CP  ∂T   ∂T  P P
[– TVβ + V]
∴ µ = – CP
β–V
TVβ V
µ = CP β – 1)
= C (Tβ … Ans.
P

Example 5.13 :
Derive a relationship for the Joule-Thomson coefficient in terms of measurable quantities
using the method of Jacobians.
Solution :
The Joule-Thomson coefficient is defined as
∂T
µ = µJT =   ... (1)
∂P H
Chemical Engineering Thermodynamics - I 5.67 Thermodynamic Properties of Pure Fluids

In Jacobian form,
[T‚ H]
µ = [P‚ H]

– [H‚ T] [H‚ T]
µ = – [H‚ P] = [H‚ P] ... (2)

The differential equation for enthalpy is


dH = T dS + V dP
This equation in the Jacobian notation is given as
[H, w] = T [S, w] + V [P, w], w is any dummy variable
Let w = T. Then
[H, T] = T [S, T] + V [P, T]
Let w = P. Then
[H, P] = T [S, P] + V [P, P] = T [S, P], since [P, P] = 0
Substituting these in Equation (2), we get
[H‚ T] T [S‚ T] + V [P‚ T]
µ = [H‚ P] = T [S‚ P]
[S‚ T] V [P‚ T]
= [S‚ P] + T [S‚ P] ... (3)

The heat capacity at constant pressure is given by


∂S [S‚ P]
CP = T   = T [T‚ P]
∂T P

CP
∴ [S, P] = T [T, P]

Substituting for [S, P] in Equation (3) gives


[S‚ T] V [P‚ T]
µ = (C /T) [T‚ P] + T (C /T) [T‚ P]
P P

T [S‚ T] V [P‚ T]
∴ µ = C [T‚ P] + C [T‚ P]
P P

We know that [x, y] = – [y, x]


T [S‚ T] V [P‚ T]
∴ µ = – C [P‚ T] – C [P‚ T]
P P
T [S‚ T] V
∴ µ = – C [P‚ T] – C ... (4)
P P
The Jacobian containing S and T indicates that it can be replaced by the Jacobian containing
P and V.
In the Jacobian notation, the Maxwell's relations are given by
[P, V] = [T, S]
[P, V] = – [S, T] ⇒ [S, T] = – [P, V]
Chemical Engineering Thermodynamics - I 5.68 Thermodynamic Properties of Pure Fluids

Substituting for [S, T], Equation (4) becomes


T – [P‚ V] V
µ = – C × [P‚ T] – C
P P

T [P‚ V] V
= C [P‚ T] – C
P P

This can be written by reversing the order of P, V and P, T as


T [V‚ P] V
µ = C [T‚ P] – C ... (5)
P P

The coefficient of volume expansion is given by


1 ∂V
β = V 
 ∂T  P

In Jacobian notation,
1 [V‚ P]
β = V [T‚ P]
[V‚ P]
∴ [T‚ P] = βV
Substituting this in Equation (5), we get
T V
µ = C (βV) – C
P P
V
∴ µ = C [Tβ β – 1] ... Ans.
P
T [V‚ P] V
OR : µ = C [T‚ P] – C
P P
T [V‚ P]
µCP = [T‚ P] – V
Express the Jacobians containing V, P and T in terms of the partial derivative to obtain
∂V
µCP = T   – V
 ∂T P
Example 5.14 :
Derive the following relation between CP and CV using the method of Jacobians :
CP – CV = β 2 VT/κ
Solution :
The heat capacity at constant pressure is given by
∂S [S‚ P]
CP = T   = T [T‚ P] ... (1)
 
∂T
P
The heat capacity at constant volume is given by
∂S [S‚ V]
CV = T   = T [T‚ V] ... (2)
∂T V
Consider the variables S, T, P and V.
Chemical Engineering Thermodynamics - I 5.69 Thermodynamic Properties of Pure Fluids

The cyclic type relation among these variables is


[S, T] [P, V] + [T, P] [S, V] + [P, S] [T, V] = 0
or [S, P] [T, V] – [S, V] [T, P] = [S, T] [P, V] ... (3)
Dividing Equation (3) by [T, P] [T, V], we get
[S‚ P] [T‚ V] [S‚ V] [T‚ P] [S‚ T] [P‚ V]
[T‚ P] [T‚ V] – [T‚ P] [T‚ V] = [T‚ P] [T‚ V]
[S‚ P] [S‚ V] [S‚ T] [P‚ V]
[T‚ P] – [T‚ V] = [T‚ P] [T‚ V] … (4)

Combining Equations (1), (2) and (4), we get


CP CV [S‚ T] [P‚ V]
T – T = [T‚ P] [T‚ V] ... (5)

The Maxwell's relations in the Jacobian notation are given by


[P, V] = [T, S]
– [V, P] = – [S, T]
[S, T] = [V, P]
With this Equation (5) becomes
CP CV [V‚ P] [P‚ V]
T – T = [T‚ P] [T‚ V]
[V‚ P] [P‚ V] T [V‚ P] [V‚ P]
CP – CV = T [T‚ P] [T‚ V] = – [T‚ P] [T‚ V] ... (6)

We know that,
1 ∂V 1 [V‚ P]
β = V   = V [T‚ P]
 ∂T  P

1 [V‚ P] [V‚ P]
β 2 = V2 [T‚ P] [T‚ P]

1 ∂V 1 [V‚ T]
κ = – V   = – V [P‚ T]
 ∂P T
1 [V‚ P] [V‚ P]
β2 V2 [T‚ P] [T‚ P]
∴ =
κ – 1  [V‚ T]
 
V [P‚ T]
– V [V‚ P] [V‚ P] [P‚ T]
= V2 [T‚ P] [T‚ P] [V‚ T]

= – V – [P‚ T] [T‚ P] [V‚ T]


1 [V‚ P] [V‚ P] [P‚ T]
 
1 [V‚ P] [V‚ P]
= V [T‚ P] [V‚ T]
Chemical Engineering Thermodynamics - I 5.70 Thermodynamic Properties of Pure Fluids

β2 1 [V‚ P] [V‚ P]
= – V [T‚ P] [T‚ V] , since [V, T] = – [T, V]
κ
[V‚ P] [V‚ P] – β 2V
∴ [T‚ P] [T‚ V] = κ ... (7)

Substituting for the Jacobians of Equation (6) for Equation (7), Equation (6) becomes
– β 2V
CP – CV = – T  
 κ 
β 2VT
∴ CP – CV = ... Ans.
κ
Example 5.15 :
Using the method of Jacobians, show that
Tβ 
dU = CV dT +  – P dV
κ 
Solution :
The required relation is given in terms of the independent variables T and V. Hence, consider
U = f (T, V)
Then the total differential of U is
∂U
dU =   dT +   dV
∂U
 ∂T  V
∂V T

Express the partial derivatives in the above equation in the Jacobian notation to get
[U‚ V] [U‚ T]
dU = [T‚ V] dT + [V‚ T] dV ... (1)
We know dU = T dS – P dV
This differential equation for U in the Jacobian notation is given as
[U, w] = T [S, w] – P [V, w] … w is any dummy variable.
Let w = V. Then
[U, V] = T [S, V] – P [V, V]
[U, V] = T [S, V], since [V, V] = 0 … [X, X] = 0
Let w = T. Then
[U, T] = T [S, T] – P [V, T]
Substituting for [U, V] and [U, T], Equation (1) becomes
T [S‚ V] T [S‚ T] – P [V‚ T]
dU = [T‚ V] dT + [V‚ T] dV
T [S‚ V] T [S‚ T]
dU = [T‚ V] dT + [V‚ T] dV – P dV ... (2)
The combination of S and V in one Jacobian indicates that it is related to CV.
∂S
CV = T  
T [S‚ V]
... CV = [T‚ V] ... (3)
∂T V
Chemical Engineering Thermodynamics - I 5.71 Thermodynamic Properties of Pure Fluids

In the Jacobian notation, the Maxwell's equations are given by


[P‚ V] = [T‚ S]
or [V, P] = [S, T] ... (4)
T [S‚ V]
Substituting for [T‚ V] from Equation (3) and for [S, T] from Equation (4) in Equation (2),
we get
T [V‚ P]
dU = CV dT + [V‚ T] dV – P dV ... (5)

The Jacobians containing P, V and T in the above equation indicate that they are related to
β and κ.
1 ∂V 1 [V‚ P]
β = V  , β = V [T‚ P] ... in Jacobian notation
 ∂T P
1 ∂V 1 [V‚ T] 1 [V‚ T]
κ = – V   , κ = – V [P‚ T] = V [T‚ P] ... in Jacobian notation
 ∂P  T

1 [V‚ P]
β V [T‚ P] [V‚ P]
= 1 [V‚ T] = [V‚ T]
κ
V [T‚ P]
With this, Equation (5) becomes

dU = CV dT + dV – P dV
κ
Tβ
dU = CV dT + 
β 
– P dV ... Ans.
κ 
Example 5.16 :
Derive the following relation through the partial derivatives method.
∂U = C – TVβ 2
 ∂T 
  V
P
κ
Solution :
We know that dU = T dS – P dV ... differential equation for U ... (1)
Consider S = f (T, P)
Then the total differential of S is
∂S ∂S
dS =   dT +   dP ... (2)
∂T ∂P
P T

∂S
We have, CP = T  
∂T P

∂S = CP
∴ ∂T
 P T
Chemical Engineering Thermodynamics - I 5.72 Thermodynamic Properties of Pure Fluids

∂S ∂V
We have, –   =   … Maxwell's equation
∂PT  ∂T P
∂S = – ∂V
∴ ∂P  ∂T 
 T  P
Substituting for the partial derivatives in Equation (2), we get
CP ∂V
dS = T dT –   dP
 ∂T  P

∂V
∴ T dS = CP dT – T   dP ... (3)
 ∂T  P

Substituting Equation (3) in Equation (1), we obtain


∂V
dU = CP dT – T   dP – P dV
 ∂T  P

Dividing by dT, holding V constant, the above equation becomes


∂U = C – T ∂V ∂P – 0
 ∂T   ∂T  ∂T
 V P
 P  V
∂U = C – T ∂V ∂P
 ∂T   ∂T  ∂T ... (4)
 V P
 P  V
We know that,
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Cyclic relation among P, V and T
     
V P T

∂P = – 1
∂T
 V   ∂V
∂T
∂V  ∂P 
 P  T
∂V
 ∂T 
 P 1 ∂y
= – as =  
∂V ∂x ∂x
 ∂P  ∂y z
 T  z
∂P
Substituting for   in Equation (4) gives
∂T V

∂V 2
T   
∂U = C +  ∂T  
P
 ∂T  ... (5)
 V P
∂V
 ∂P 
 T
Chemical Engineering Thermodynamics - I 5.73 Thermodynamic Properties of Pure Fluids

The partial derivative terms in Equation (5) are related to β and κ. We know that
1 ∂V ∂V
β = V  ∴   = βV ... (6)
 ∂T 
P
 ∂T 
P

1 ∂V ∂V
κ = –V  ∴   = – κV ... (7)
 ∂P T  ∂P 
T

β = Coefficient of volume expansion


κ = Coefficient of compressibility
Substituting Equations (6) and (7) in Equation (5), we get
∂U = C + T [βV]2
 ∂T 
  V
P
(– κV)

∂U = C – TVβ
β2
∴  ∂T  ... Ans.
 V P
κ
Example 5.17 :
Determine β and κ for a van der Waals gas.
Solution :
The coefficient of volume expansion is given by
1 ∂V
β = V  ... (1)
 ∂T  P

The van der Waals equation of state is given by


P + a  (V – b) = RT
 V 2
where a and b are van der Waals constants.
Differentiating the van der Waals equation with respect to T at constant P, we get
P + a  ∂V + (V – b) – 2a ∂V  = R
 V2  ∂T   V3  ∂T  
P   P

P + a  ∂V – 2a (V – b) ∂V = R
 V2  ∂T P V3  ∂T P
∂V P + a  – 2a (V – b) = R
 ∂T   V2 
 P V3

∂V = R
∴  ∂T 
 P P + a  – 2a (V – b)
 V 2 V3
Multiplying both the numerator and denominator by (V – b), we get
∂V = R (V – b)
 ∂T 
 P P +  (V – b) – 2a (V – b)2
a
 V 2 V3
Chemical Engineering Thermodynamics - I 5.74 Thermodynamic Properties of Pure Fluids

Substituting RT for P + V2 (V – b) in the denominator of the above equation, we get


a
 
∂V = R (V – b)
 ∂T 
 P RT – 2a (V – b)2
V3
∂V = R (V – b)
 ∂T 
 P RTV 3 – 2a (V – b)2

V3
∂V = RV3 (V – b)
 ∂T  ... (2)
 P RTV3 – 2a (V – b)2
Substituting Equation (2) in Equation (1), we get

β = V RTV3 – 2a (V – b)2
1 RV3 (V – b)
 
RV2 (V – b)
∴ β = RTV3 – 2a (V – b)2 ... Ans. (i)

The coefficient of compressibility is given by


1 ∂V
κ = –V  ... (3)
 ∂P T

P + a  (V – b) = RT
 V 2
Differentiating the above equation with respect to V at constant T,

P + a  (1) + (V – b)  ∂P  – 2a  = 0
 V 2 ∂V V3
 T 

P + a  + (V – b)  ∂P  – 2a  = 0
 V 2 ∂V V3
 T 
 ∂P 
(V – b)   – V3 (V – b) + P + V2 = 0
2a a
 
∂V
T
 

 ∂P  2a (V – b) 
– P + V 2
a
(V – b)   =
 
∂V V 3
 
T

Multiplying and dividing the last term on the RHS by (V – b),

P + a  (V – b)
 
∂P 2a (V – b)  V 2
(V – b)   = –
∂V T
V3 (V – b)
Chemical Engineering Thermodynamics - I 5.75 Thermodynamic Properties of Pure Fluids

Since P + V2 (V – b) = RT,


a
 
 ∂P  2a (V – b) RT
(V – b)   = – (V – b)
 
∂V
T
V 3

 ∂P  = 2a (V – b) – RT
∂V
 T V3 (V – b) (V – b)2
 ∂P  = 2a – RT
∂V
 T V3 (V – b)2
 ∂P  = 2a (V – b)2 – RTV3
∂V
 T V3 (V – b)2

 ∂P  = 1
We know that, ∂V ... Reciprocal relation
 T ∂V
 ∂P 
 T
Using this reciprocal relation, we can write
1 2a (V – b)2 – RTV3
=
∂V V3 (V – b)2
 ∂P 
  T

∂V = V3 (V – b)2
∴  ∂P  ... (4)
 T 2a (V – b)2 – RTV3
Substituting Equation (4) in Equation (3) gives

κ = – V 2a (V – b)2 – RTV3


1 V3 (V – b)2
 
– [V2 (V – b)2]
κ =
– [RTV3 – 2a (V – b)2]
V2 (V – b)2
∴ κ = RTV3 – 2a (V – b)2 ... Ans.

Example 5.18 :
Using the method of Jacobians, show that
∂H = T – 1
 ∂S 
 T β
Solution :
We know that : dH = T dS + V dP ... (1)
In Jacobian notation, it is written as
[H, w] = T [S, w] + V [P, w]
Let w = T. Then
Chemical Engineering Thermodynamics - I 5.76 Thermodynamic Properties of Pure Fluids

[H, T] = T [S, T] + V [P, T]


∂H
Express   in the Jacobian form.
 ∂S  T

∂H = [H‚ T]
 ∂S 
 T [S‚ T]

Substituting for [H, T] gives


∂H [H‚ T] T [S‚ T] + V [P‚ T]
 ∂S  = [S‚ T] =
 T [S‚ T]

T [S‚ T] V [P‚ T]
= [S‚ T] + [S‚ T]

∂H = T + V [P‚ T]
 ∂S  ... (2)
 T [S‚ T]
The Maxwell’s equations in the Jacobian form are given by
[P, V] = [T, S]
This can be written as
[V, P] = [S, T]
[S, T] = [V, P]
Substituting this in Equation (2) gives
∂H = T + V [P‚ T] = T – V [T‚ P] , since [x, y] = – [y, x]
 ∂S  ... (3)
 T [V‚ P] [V‚ P]

1 ∂V
We have, β = V 
 ∂T P

In terms of Jacobians, it becomes


1 [V‚ P]
β = V [T‚ P]
Rearranging this, we get
V [T‚ P] 1
[V‚ P] = β ... (4)

Substituting Equation (4) into Equation (3) gives


∂H = T – 1
 ∂S  ... Ans.
 T β
Example 5.19 :
Using the method of Jacobians, show that
∂H = T 1 + Vβ 
 ∂S   C κ
  V
 V 
Chemical Engineering Thermodynamics - I 5.77 Thermodynamic Properties of Pure Fluids

Solution :
We have : dH = T dS + V dP
This equation in Jacobian notation is written as
[H, w] = T [S, w] + V [P, w]
Let w = V. Then
[H, V] = T [S, V] + V [P, V] ... (1)
∂H
Express   in the Jacobian form.
 ∂S  V

∂H = [H‚ V]
 ∂S  ... (2)
 V [S‚ V]

Combining Equations (1) and (2) gives


∂H = T [S‚ V] + V [P‚ V]
 ∂S 
 V [S‚ V]

T [S‚ V] V [P‚ V]
= [S‚ V] + [S‚ V]

∂H = T + V [P‚ V]
 ∂S  ... (2)
 V [S‚ V]

1 ∂V 1 ∂V
We have, β = V  κ = –V 
 ∂T P
 ∂P  T

1 [V‚ P] 1 [V‚ T]
β = V [T‚ P] κ = – V [P‚ T]

∂S
CV = T  
∂T V

[S‚ V]
CV = T [T‚ V]

β 1 [V‚ P] 1 [T‚ V] [P‚ T]


∴ = V [T‚ P] × T [S‚ V] × (– V) [V‚ T]
CV κ
1 [V‚ P] [T‚ V] [P‚ T]
= – T [T‚ P] [S‚ V] [V‚ T]
1 [V‚ P] [T‚ V] [T‚ P] 1 [V‚ P]
= – T [T‚ P] [S‚ V] [T‚ V] = – T [T‚ P]
[Since [P, T] = – [T, P] and [V, T] = – [T, V]
β 1 [V‚ P] 1 [P‚ V]
= – T [T‚ P] = T [T‚ P]
CV κ
[P‚ V]  β 
∴ [T‚ P] = T CV κ ... (3)
Chemical Engineering Thermodynamics - I 5.78 Thermodynamic Properties of Pure Fluids

Substituting Equation (3) into Equation (1), we get


∂H = T + V·T  β 
 ∂S  C κ
  V
 V 
∂H = T 1 + Vββ
∴  ∂S   C κ ... Ans.
 V  V 

Example 5.20 :
Using the method of Jacobians, show that
∂P = β
∂T
  κ
V

Solution :
The coefficient of compressibility is given by
1 ∂V
κ = –V 
 ∂P  T

In Jacobian form, it becomes


1 [V‚ T]
κ = – V [P‚ T]
The coefficient of volume expansion is given by
1 ∂V
β = V 
 ∂T  P
In Jacobian form, it becomes
1 [V‚ P]
β = V [T‚ P]
β 1 [V‚ P] 1
= V [T‚ P]
κ –  [V‚ T]
1

V [P‚ T]
β [V‚ P] [P‚ T]
= – [T‚ P] [V‚ T]
κ
β  [P‚ V] [P‚ T]  [P‚ V] [P‚ T]
= – – [T‚ P] [V‚ T]  = [V‚ T] [T‚ P]
κ  
[P‚ V] [P‚ T] [P‚ V] [P‚ T]
= – [T‚ V] × [T‚ P] = – [T‚ V] × – [P‚ T]
β [P‚ V] [P‚ T] [P‚ V]
= [T‚ V] [P‚ T] = [T‚ V] … (1)
κ
∂P = [P‚ V]
We have, ∂T … (2)
 V [T‚ V]
Combining Equations (1) and (2) gives

=  
β ∂P
... Ans.
κ ∂T V
Chemical Engineering Thermodynamics - I 5.79 Thermodynamic Properties of Pure Fluids

Example 5.21 :
Using the method of Jacobians, show that
∂H = – 1 + T ∂P
∂V ∂T
 T κ  V
Solution :
We have, dH = T dS + V dP
In Jacobian notation, the above equation is written as
[H‚ w] = T [S‚ w] + V [P‚ w]
Let w = T. Then,
[H‚ T] = T [S, T] + V [P, T] ... (1)
∂H
Express   in Jacobian form using the property of Jacobians.
∂V T

∂H = [H‚ T]
∂V ... (2)
 T [V‚ T]
Combining Equations (1) and (2) gives
∂H [H‚ T] T [S‚ T] + V [P‚ T]
∂V = [V‚ T] =
  T
[V‚ T]

∂H = T [S‚ T] + V [P‚ T]


∂V ... (3)
 T [V‚ T] [V‚ T]
In the Jacobian notation, the Maxwell's equations are given by
[P‚ V] = [T‚ S]
It can be written as
[P‚ V] [T‚ S]
[V‚ T] = [V‚ T]
[T‚ S] [S‚ T] [P‚ V] [V‚ P]
[V‚ T] = – [V‚ T] = [V‚ T] = – [V‚ T]
[S‚ T] [V‚ P]
∴ [V‚ T] = [V‚ T]
With this, Equation (3) becomes
∂H = T [V‚ P] + V [P‚ T]
∂V ... (4)
  T
[V‚ T] [V‚ T]
[V‚ P] [P‚ V] ∂P
We have, = = ∂T ... (5)
[V‚ T] [T‚ V]  V
The coefficient of compressibility (κ) is given by
1 ∂V
κ = –V 
 ∂P T
Chemical Engineering Thermodynamics - I 5.80 Thermodynamic Properties of Pure Fluids

In terms of Jacobian, it becomes


1 [V‚ T]
κ = – V [P‚ T]

[P‚ T] 1
∴ [V‚ T] = – κV ... (6)

Combining Equations (4), (5) and (6) gives


∂H = T ∂P + V – 1 
∂V ∂T
 T  V  κV
∂H = – 1 + T ∂P
 ∂V  ∂T ... Ans.
 T κ  V
Example 5.22 :
Using the method of Jacobians, show that
∂U = T ∂P – P
∂V ∂T
   
T V

Solution :
We have : dU = T dS – P dV
In Jacobian form, it becomes
[U‚ w] = T [S, w] – P [V, w]
Let w = T. Then,
[U, T] = T [S, T] – P [V, T] ... (1)
∂U
Express   in Jacobian notation.
∂V T

∂U = [U‚ T]
∂V ... (2)
 T [V‚ T]
Combining Equations (1) and (2) gives
∂U = T [S‚ T] – P [V‚ T]
∂V
  T
[V‚ T]
[S‚ T] [V‚ T]
= T [V‚ T] – P [V‚ T]
[S‚ T]
= T [V‚ T] – P
We know, [P, V] = [T, S] ... Maxwell’s equation
i.e., [P, V] = – [S, T]
∂U [P‚ V]
∴ ∂V = – T [V‚ T] – P
  T
[P‚ V] [P‚ V]
= – T – [T‚ V] – P = T [T‚ V] – P … (3)
Chemical Engineering Thermodynamics - I 5.81 Thermodynamic Properties of Pure Fluids

∂P = [P‚ V]
We have, ∂T ... (4)
 V [T‚ V]

[P‚ V]
Substituting for [T‚ V] , from Equation (4), Equation (3) becomes

∂U = T ∂P – P
∂V ∂T ... Ans.
 T  V
Example 5.23 :
Using the method of Jacobians, show that
∂G = – 1
∂V
 T κ
Solution :
We have : dG = – S dT + V dP ... (1)
Equation (1) in terms of Jacobians becomes
[G, w] = – S [T, w] + V [P, w]
For w = T : [G, T] = – S [T, T] + V [P, T]
[G, T] = 0 + V [P, T] … as [T, T] = 0
[G, T] = V [P, T] ... (2)
∂G
Express   in Jacobian form
∂V T

∂G = [G‚ T]
∂V ... (3)
 T [V‚ T]
Substituting for [G, T] from Equation (2), Equation (3) becomes
∂G = [G‚ T] = V [P‚ T]
∂V
 T [V‚ T] [V‚ T]

Eliminate the Jacobian in terms of the partial derivative.


∂G [P‚ T]  ∂P 
∂V = V [V‚ T] = V ∂V ... (4)
 T  T
The coefficient of compressibility is given by
1 ∂V
κ = –V 
 ∂P  T

1 ∂y 1
∴ κ = – , as   =
 ∂P   z  
∂x ∂x
V  ∂y
∂V T
  z
Chemical Engineering Thermodynamics - I 5.82 Thermodynamic Properties of Pure Fluids

 ∂P  1
∴ V  = – ... (5)
∂V κ
T

Combining Equations (4) and (5), we get


∂G = – 1
 ∂V  ... Ans.
  T
κ
Example 5.24 :
∂A
Evaluate   in terms of measurable quantities using the method of Jacobians.
∂G T

Solution :
∂A = [A‚ T]
∂G
 T [G‚ T]
We know that dA = – S dT – P dV ... Differential equation for A
and dG = – S dT + V dP ... Differential equation for G
These in the Jacobian notation are given as
[A, w] = – S [T, w] – P [V, w]
[G, w] = – S [T, w] + V [P, w]
Let w = T. Then,
[A, T] = – S [T, T] – P [V, T] = 0 – P [V, T] = – P [V, T]
[G, T] = – S [T, T] + V [P, T] = V [P, T], since [x, x] = 0
[A‚ T] – P [V‚ T]
∴ [G‚ T] = V [P‚ T] ... (1)
The Jacobians containing V, T and P as given in the above equation are related to κ.
We have,
1 ∂V 1 [V‚ T]
κ = – V   = – V [P‚ T]
 
∂P
T
[V‚ T]
∴ [P‚ T] = – κV ... (2)
The Jacobians containing V, P and T can be eliminated by substituting Equation (2) in
Equation (1). Substituting Equation (2) in Equation (1), we get
[A‚ T] P
[G‚ T] = – V (– κV)
[A‚ T]
Therefore, κ
[G‚ T] = Pκ ... Ans.
Example 5.25 :
Using the method of Jacobians, show that
T dS = CP dT – β VT dP
Solution :
Consider S = f (T, P)
Chemical Engineering Thermodynamics - I 5.83 Thermodynamic Properties of Pure Fluids

The total differential of S is


∂S ∂S
dS =   dT +   dP
∂T P
∂P T

In the Jacobian notation,


[S‚ P] [S‚ T]
dS = [T‚ P] dT + [P‚ T] dP ... (1)

The Jacobians containing S, P and T are related to CP. The heat capacity at constant pressure
is given by
∂S
CP = T  
∂T P

[S‚ P]
CP = T [T‚ P]

[S‚ P] CP
[T‚ P] = T ... (2)

Substituting Equation (2) in Equation (1), we get


CP [S‚ T]
dS = T dT + [P‚ T] dP ... (3)

In the Jacobian notation, the Maxwell's relations are given by


[P, V] = [T, S] = – [S, T]
With this, Equation (3) becomes
CP [P‚ V]
dS = T dT – [P‚ T] dP

CP [V‚ P]
dS = T dT – [T‚ P] dP , using [x, y] = – [y, x] ... (4)
The Jacobians containing V, P and T are related to β.
The coefficient of volume expansion is given by
1 ∂V 1 [V‚ P]
β = V   = V [T‚ P]
 ∂T  P
[V‚ P]
[T‚ P] = Vβ ... (5)
Combining Equations (4) and (5) gives
CP
dS = T dT – Vβ dP
∴ T dS = CP dT – β VT dP ... Ans.
Example 5.26 :
Prove that CP is independent of pressure for a gas obeying the relation P (V – b) = RT.
Solution :
Here we have to show that CP does not change with P at constant T.
Chemical Engineering Thermodynamics - I 5.84 Thermodynamic Properties of Pure Fluids

The heat capacity at constant pressure is given by


∂S
CP = T  
∂T P

∂S = CP
∂T ... (1)
 P T

Differentiate Equation (1) with respect to P keeping T constant. Doing this, we get
∂2S ∂CP/T 1 ∂CP
=   = T  ... (2)
∂P ∂T  ∂P T  ∂P T
The Maxwell's relation in terms of S, P related to the measurable quantities V, T is given by
∂S ∂V
–  =  
 T  ∂T P
∂P

∂S = – ∂V
∂P  ∂T  ... (3)
 T  P
Differentiate Equation (3) with respect to T keeping P constant. This results in
∂2S ∂2V
= –  2 ... (4)
∂T ∂P  ∂T P
∂2z ∂2z
We know that, =
∂x ∂y ∂y ∂x
∂2S ∂2S
∴ =
∂P ∂T ∂T ∂P
Substituting for these terms from Equations (2) and (4) gives
1 ∂CP ∂2V
 
T  ∂P  = –  ∂T2 
T
  P

∂CP ∂2V
 ∂P  = – T  2  ... (5)
 T  ∂T P

∂2V
Hence, we have to find  2  and this has to be zero for CP to be independent of pressure.
 ∂T  P

The equation of state is


P (V – b) = RT
RT
(V – b) = P

V = P T + b
R
 
Chemical Engineering Thermodynamics - I 5.85 Thermodynamic Properties of Pure Fluids

Differentiating this equation w.r.t. T at constant P, we get


∂V = R + 0 = R
 ∂T 
 P P P

Again differentiating the above equation w.r.t. T at constant P, we get


∂2V = 0
 ∂T2 
 P
Substituting this in Equation (5),
∂CP
  = – T (0) = 0 ... (6)
 ∂P T
It is evident from Equation (6) that CP is independent of pressure for the given gas obeying
the given equation of state. … Ans.
Example 5.27 :
Derive the following relation using the method of Jacobians :
∂U = Tβ – P
 ∂V 
 T κ
Solution :
The differential of U is given by
dU = T dS – P dV
In the Jacobian notation,
[U, w] = T [S, w] – P [V, w]
Let w = T. Then,
[U, T] = T [S, T] – P [V, T]
Dividing by [V, T], we get
[U‚ T] [S‚ T] [V‚ T]
[V‚ T] = T [V‚ T] – P [V‚ T]
[U‚ T] [S‚ T]
[V‚ T] = T [V‚ T] – P
In the Jacobian notation, the Maxwell's relations are given by
[P, V] = [T, S]
[P, V] = – [S, T]
[U‚ T] [P‚ V]
∴ [V‚ T] = – T [[V‚ T] – P
Replace the Jacobians containing U, V and T by the partial derivative term to get
∂U = – T [P‚ V] – P
∂V ... (1)
  T
[V‚ T]
Chemical Engineering Thermodynamics - I 5.86 Thermodynamic Properties of Pure Fluids

The Jacobians containing P, V and T present in the above equation are related to β and κ
as mentioned below :
1 ∂V 1 [V‚ P]
β = V   = V [T‚ P]
 ∂T P
1 ∂V 1 [V‚ T]
κ = – V   = – V [P‚ T]
 ∂P T
β – [V‚ P] [P‚ T] [P‚ V] [P‚ T] – [P‚ V]
∴ = [T‚ P] [V‚ T] = – [P‚ T] [V‚ T] = [V‚ T]
κ
Substituting β/κ for the Jacobians, Equation (1) becomes
∂U = Tββ
∂V –P ... Ans.
 T κ

Example 5.28 :
Derive the following relation using the method of Jacobians.
∂H
 ∂P  = V (1 – Tβ)
 T
Solution :
We know that dH = T dS + V dP
In the Jacobian notation,
[H‚ w] = T [S‚ w] + V [P‚ w]
Let w = T. Then
[H‚ T] = T [S‚ T] + V [P‚ T]
Divide by [P‚ T]
[H‚ T] [S‚ T] [P‚ T]
∴ [P‚ T] = T [P‚ T] + V [P‚ T]
[H‚ T] [S‚ T]
[P‚ T] = T [P‚ T] + V
Replace the Jacobians containing H, P and T in terms of partial derivative to get
∂H = T [S‚ T] + V
 ∂P 
  T
[P‚ T]
The Maxwell's relations in terms of the Jacobians are given by
[P, V] = [T, S] = – [S, T]
∂H = – T [P‚ V] + V
 ∂P 
  T
[P‚ T]
This can be written as
∂H = – T [V‚ P] + V
 ∂P  ... (1)
  T
[T‚ P]
Chemical Engineering Thermodynamics - I 5.87 Thermodynamic Properties of Pure Fluids

The Jacobians containing V, P and T is related to β.


β = Coefficient of volume expansion
1 ∂V
= V 
 ∂T P
In the Jacobian notation,
1 [V‚ P]
β = V [T‚ P]

[V‚ P]
∴ [T‚ P] = βV ... (2)

Combining Equations (1) and (2) gives


∂H = – TβV + V
 ∂P 
 T
∂H = V (1 – Tβ
 ∂P  β) ... Ans.
 T
Example 5.29 :
Prove that U is a function of temperature alone for a gas obeying the relation
P (V – b) = RT.
Solution :
We have to prove that U depends on temperature alone. That is, we have to prove that U does
not change with P or V at constant T.
The differential equation for U is
dU = T dS – P dV ... (1)
We have to express U in terms of the measurable quantities such as P, V and T.
Let S = f (P, T)
The total differential of S is
∂S ∂S
dS =   dP +   dT ... (2)
∂P T
∂T P

We know that,
∂S
CP = T  
∂T P

∂S = CP
∴ ∂T
 P T

∂S = – ∂V … Maxwell's equation


and ∂P  ∂T 
 T  P
Chemical Engineering Thermodynamics - I 5.88 Thermodynamic Properties of Pure Fluids

Substituting for the partial derivative terms, Equation (2) becomes


CP ∂V
dS = T dT –   dP
 ∂T  P

∂V
T dS = CP dT – T   dP
 ∂T  P

Substituting for T dS in Equation (1) gives


∂V
dU = CP dT – T   dP – P dV
 ∂T  P

Dividing by dV holding T constant, we obtain


∂U = 0 – T ∂V  ∂P  – P
∂V  ∂T  ∂V
  T
    P T

∂U = – T ∂V  ∂P  – P
∂V  ∂T  ∂V ... (3)
 T  P  T
We have :
∂P  ∂T  ∂V = – 1 … cyclic relation among P, T and V.
∂T ∂V  ∂P 
     
V P T

∂P = –1 ∂V  ∂P 
∂T =–    ... Using reciprocal relation
 V    
∂T ∂V  ∂T P ∂VT
∂V  ∂P 
 P  T
∂V  ∂P 
Substituting for –     from the above equation in Equation (3) gives
 ∂T  ∂V
P T

∂U = T ∂P – P
∂V ∂T ... (4)
 T  V
The equation of state is
P (V – b) = RT
P = (V – b) = (V – b) T
RT R
 
Differentiating this equation w.r.t. T at constant V,
∂P = R (1) = R
∂T
  V–b
V
V–b

∂P = R
∂T ... (5)
 V V–b
Substituting Equation (5) in Equation (4), we get
∂U = RT – P
∂V
  (V – b)
T
Chemical Engineering Thermodynamics - I 5.89 Thermodynamic Properties of Pure Fluids

RT
But (V – b) = P
∂U = P – P = 0
∴ ∂V
 T
∂U = 0
∴ ∂V ... (6)
 T
This means U does not depend on V.
We know that,
∂U = ∂U  ∂P 
∂V     ... Chain rule of partial differentiation
 T  ∂P T ∂VT
 ∂P 
It is clear from the equation of state that   ≠ 0.
∂V T

∂U = 0
We have, ∂V
 T
∂U = 0 = ∂U  ∂P  but  ∂P  ≠ 0
Therefore, ∂V  ∂P  ∂V ∂V
 T  T  T  T
∂U = 0
Therefore,  ∂P  ... (7)
 T
This means that U does not depend on P.
Equations (6) and (7) indicate that U depends only on T.
Example 5.30 :
RT2 ∂Z
Prove that µ = PC  
P ∂T
P

where µ is the Joule-Thomson coefficient and Z is the compressibility factor.


Solution :
The Joule-Thomson coefficient is defined as
∂T
µ = µJT =  
∂P H

To express the Joule-Thomson coefficient of a gas in terms of the measurable quantities,


consider the relation f (H, T, P) = 0 and write the cyclic relation as
∂H ∂T  ∂P  = – 1
 ∂T  ∂P ∂H ... Cyclic relation among H, T and P
 P  H  T
Chemical Engineering Thermodynamics - I 5.90 Thermodynamic Properties of Pure Fluids

∂T = – 1
∴ ∂P
 H ∂H  ∂P 
 ∂T  ∂H
 P  T

∂H
 ∂P 
∂T = µ = –  T 1 ∂y
∴ ∂P as =  
 H ∂H ∂x ∂xz
 ∂T  ∂y
 P  z

We know that :
∂H
CP =  
 ∂T P
∂H
Substituting for   in the above relation of µ, we get
 ∂T  P

∂H
 ∂P 
 T
µ = – C ... (1)
P

We have, dH = T dS + V dP
Dividing by dP holding T constant, Equation (1) becomes
∂H = T ∂S + V
 ∂P  ∂P ... (2)
   
T T

Substituting the Maxwell's relation given by


∂S ∂V ∂S ∂V
–  =   ⇒   = – 
∂P  ∂T 
T P
∂P  ∂T  T P

in Equation (2), we obtain


∂H = – T ∂V + V
 ∂P   ∂T  ... (3)
   
T P

Substituting Equation (3) in Equation (1), we get


 ∂V 
– – T   + V
  ∂T   P
µ = CP
∂V
µCP = T   – V ... (4)
 ∂T  P

The compressibility factor is given by


PV
Z = RT

Z = R T
P V
 
Chemical Engineering Thermodynamics - I 5.91 Thermodynamic Properties of Pure Fluids

Differentiating the above relation with respect to T at constant P gives


 ∂V 
T  ∂T  – V
∂Z = P  P 
∂T  
 P R T 2

∂Z = P T ∂V – V
∂T  
 P RT2   ∂T P 

RT2 ∂Z ∂V
∴ P ∂T = T  ∂T  – V ... (5)
P P

The R.H.S. of Equation (4) is identical to that of Equation (5). Hence, combining Equations
(4) and (5) gives
RT2 ∂Z
µCP = P ∂T
P

RT2 ∂Z
∴ µ = PC   ... Ans.
P ∂T
P

Example 5.31 :
Derive the following relation using the partial derivatives method.
∂U = Tβ – P
∂V
 T κ

Solution :
We know that,
dU = T dS – P dV ... (1)
Dividing by dV holding T constant gives
∂U  ∂S 
∂V = T ∂V – P ... (2)
 T  T
We know that,
 ∂S  = ∂P
∂V   ... Maxwell’s relation
 T ∂TV
Substituting this Maxwell's relation in Equation (2), we get
∂U = T ∂P – P
∂V ∂T ... (3)
 T  V
We have for f (P, T, V) = 0
∂P  ∂T  ∂V = – 1
∂T ∂V  ∂P  ... Cyclic relation
 V  P  T
Chemical Engineering Thermodynamics - I 5.92 Thermodynamic Properties of Pure Fluids

∂P = – 1
∴ ∂T
 V  ∂T  ∂V
∂V  ∂P 
 P  T

∂V
 ∂T 
 P
= – ... Using reciprocal relation
∂V
 ∂P 
 T

∂P
Substituting for   in Equation (3), it becomes
∂T V

∂V
 ∂T 
∂U = – T  P
∂V –P ... (4)
 T ∂V
 ∂P 
 T
1 ∂V
We know that, β = V  ... Coefficient of volume expansion
 ∂T P

∂V = βV
∴  ∂T 
 P
1 ∂V
and κ = –V  ... Coefficient of compressibility
 ∂P T
∂V = – κV
∴  ∂P 
 T
Substituting the partial derivatives in Equation (4) in terms of the measurable quantities,
we get
∂U (βV)
∂V = – T (– κV) – P
 T
∂U = Tββ
∂V –P ... Ans.
 T κ
Extra : Use of Jacobians method :
dU = T dS – P dV ... (1)
In the Jacobian notation,
[U, w] = T [S, w] – P [V, w]
Let w = T. Then,
[U, T] = T [S, T] – P [V, T] ... (2)
Chemical Engineering Thermodynamics - I 5.93 Thermodynamic Properties of Pure Fluids

The Maxwell's relations in the Jacobian form are written as


[P, V] = [T, S] = – [S, T]
Substituting for [S, T], Equation (2) becomes
[U, T] = – T [P, V] – P [V, T]
Dividing by [V, T], we get
[U‚ T] [P‚ V] [V‚ T]
[V‚ T] = – T [V‚ T] – P [V‚ T]
[U‚ T] [P‚ V]
[V‚ T] = – T [V‚ T] – P
This can be written as
[U‚ T] [P‚ V]
[V‚ T] = T [T‚ V] – P ... (3)

The Jacobians containing P, V and T in Equation (3) indicates that they are related to
β and κ.
1 ∂V 1 [V‚ P]
β = V   = V [T‚ P]
 P
∂T

1 ∂V 1 [V‚ T]
κ = – V   = – V [P‚ T]
 ∂P T
1 [V‚ P]
β V [T‚ P]
= 1 [V‚ T]
κ
– V [P‚ T]

[V‚ P] [P‚ T] [P‚ V] [P‚ T] [P‚ V]


= – [T‚ P] [V‚ T] = [T‚ V] [P‚ T] = [T‚ V]

β
Substituting for the Jacobians containing P, V and T by and replacing the Jacobians
κ
containing U, V and T by the partial derivatives, Equation (3) becomes
∂U = Tββ
∂V –P ... Ans.
 T κ

Example 5.32 :
Determine the Joule-Thomson coefficient for a van der Waals gas.
Solution :
The Joule-Thomson coefficient is defined as
∂T
µ =  
∂P H
Chemical Engineering Thermodynamics - I 5.94 Thermodynamic Properties of Pure Fluids

To express the Joule-Thomson coefficient of a gas in terms of the measurable quantities,


consider the relation f (H, T, P) = 0 and write the cyclic relation among these quantities as
∂H ∂T  ∂P  = – 1
 ∂T  ∂P ∂H
 P  H  T
∂T = – 1
∂P
 H ∂H  ∂P 
 ∂T  ∂H
 P  T

∂H
 ∂P 
 T
= – ... Using reciprocal relation
∂H
 ∂T 
 P
The heat capacity at constant pressure is given by
∂H
CP =  
 ∂T  P

∂H
 ∂P 
∂T = µ = –  T
∴ ∂P ... (1)
 H CP

We know that dH = T dS + V dP
Dividing by dP holding T constant gives
∂H = T ∂S + V
 ∂P  ∂P ... (2)
 T  T
Substituting the Maxwell's relation given by
∂S = – ∂V
∂P  ∂T 
 T  P
in Equation (2), we get
∂H = – T ∂V + V
 ∂P   ∂T  ... (3)
 T  P
Substituting Equation (3) in Equation (1), we get
∂V
T  –V
 ∂T 
P
µ = CP
∂V
∴ µCP = T   – V ... (4)
 ∂T  P
Chemical Engineering Thermodynamics - I 5.95 Thermodynamic Properties of Pure Fluids

∂V
Let us evaluate   for the van der Waals gas.
 ∂T  P

The van der Waals equation of state is


P + a  (V – b) = RT ... (5)
 V 2
Differentiating Equation (5) w.r.t. T at constant P, we get
P + a  ∂V + (V – b) 0 – 2a ∂V  = R
 V2  ∂T   V3  ∂T  
P   P

Multiplying this equation throughout by (V – b), we obtain


P + a  (V – b) ∂V + (V – b)2 – 2a ∂V  = R (V – b)
 V 2  ∂T   V3  ∂T  
P   P

With P + V2 (V – b) = RT, the above equation becomes


a
 
∂V 2a (V – b)2 ∂V = R (V – b)
RT   –  ∂T 
 ∂T  P
V3   P

∂V RT – 2a (V – b)2  = R (V – b)


 ∂T   
 P V3

  ∂V
[RTV3 – 2a (V – b)2]  ∂T  = RV3 (V – b)
 P
∂V = RV3 (V – b)
∴  ∂T  ... (6)
 P RTV3 – 2a (V – b)2
∂V
Substituting for   in Equation (4) gives
 ∂T  P

TRV3 (V – b)
µCP = RTV3 – 2a (V – b)2 – V

Taking V common from the terms on R.H.S., we get

µCP = V RTV3 – 2a (V – b)2 – 1


RTV2 (V – b)
 

µCP = V 
RTV2 (V – b) – [RTV3 – 2a (V – b)2]
 RTV3 – 2a (V – b)2 

µCP = V  
RTV3 – RTV2b– RTV3
+ 2a (V – b)2
 RTV3 – 2a (V – b)2 

µCP = V  RTV3 – 2a (V – b)2 


2a (V – b)2 – RTV2 b
 

µ = C  RTV3 – 2a (V – b)2 
V 2a (V – b)2 – RTV2 b
∴ ... Ans.
P  
Chemical Engineering Thermodynamics - I 5.96 Thermodynamic Properties of Pure Fluids

Example 5.33 :
γ
Prove that P (V – b) is constant for a reversible adiabatic process for a gas obeying the
relation P (V – b) RT = C.
Solution :
A reversible adiabatic process is called as an isentropic process (a process with constant S).
Consider S = f (P, V)
The total differential of S is
∂S  ∂S 
dS =   dP +   dV
∂P V
∂VP
At constant S, this becomes
∂S  ∂S 
∂P dP + ∂V dV = 0
  V
  P
Dividing by dV holding S constant, we obtain
∂S  ∂P  +  ∂S  = 0
∂P ∂V ∂V
    V
 
S P

 ∂S 
∂V
 ∂P  = –  P
∴ ∂V ... (1)
 S ∂S
∂P
 V
The chain rule of partial differentiation is
∂x = ∂z ∂x
∂a     
 y ∂xy ∂a y
 ∂S  = ∂S  ∂T 
∴ ∂V     ... (2)
 P ∂TP ∂VP
∂S ∂S ∂T
∂P = ∂T ∂P ... (3)
 V  V  V
Substituting Equations (2) and (3) in Equation (1), we get
∂S  ∂T 
∂T ∂V
 ∂P  = –  P  P
∂V ... (4)
 S ∂S ∂T
∂T ∂P
 V  V
We know that,
∂S ∂S CP
CP = T   ∴   = T ... (5)
∂T P
 
∂T
P

∂S ∂S CV
CV = T   ∴   = T ... (6)
∂T V
 
∂T
V
Chemical Engineering Thermodynamics - I 5.97 Thermodynamic Properties of Pure Fluids

Substituting for partial derivatives from Equations (5) and (6), Equation (4) becomes
CP  ∂T   ∂T 
– T   ∂V
 ∂P  = ∂V P CP  P
∂V = –C
 S CV ∂T V ∂T
T ∂P  ∂P
 
V V

CP
We know that, CV = γ. With this the above equation becomes
 ∂T 
∂V
 ∂P  = – γ  P
∂V ... (7)
 S ∂T
∂P
 V
The equation of state is
P (V – b) = RT

P = V – b = (V – b) T
RT R
 
Differentiating with respect to T at constant V yields
∂P =  R  × 1 = R
∂T V – b
 V V–b

P (V – b) = RT

V – b = P T
R
 

V = P T + b
R
 
Differentiating with respect to T at constant P gives
∂V = R (1) + 0 = R
 ∂T 
 P P P

∂V R ∂V 1
 ∂T  = P ...   = ... Reciprocal relation
 P  P  ∂T 
∂T
∂V
  P

∂V 1 R
 ∂T  = ∂T = P
 P  
∂V
 P

 ∂T  = P
∴ ∂V ... (8)
 P R
Chemical Engineering Thermodynamics - I 5.98 Thermodynamic Properties of Pure Fluids

∂P = 1 = R
∂T
 V ∂T V–b
∂P
 V
∂T = (V – b)
∴ ∂P ... (9)
 V R
Substituting Equations (8) and (9) in Equation (7), we get
 ∂P  = – γ (P/R) = – γP
∂V
  (V – b)/R
S
V–b
This can be written as
dP – γ dV
P = (V – b)
d ln P = – γ d ln (V – b)
d ln P = – d ln (V – b)γ
d ln P + d ln (V – b)γ = 0
d ln P (V – b)γ = 0
i.e. P (V – b)γ = C … Ans.
where C is a constant.
dP dV
OR : P = –γV–b
ln P = – γ ln (V – b) + C1
ln P + γ ln (V – b) = C1
ln P + ln (V – b)γ = C1
ln P (V – b)γ = C1
∴ P (V – b)γ = C
where C is a constant.
Example 5.34 :
A pure gas flowing through a well insulated horizontal pipe at high pressure is throttled to a
slightly low pressure. The gas obeys the relation P (V – b) = RT, where b is a positive constant.
Does the temperature of the gas rise or fall by throttling ?
Solution :
We know that :
For µ = µJT > 0 ... throttling results in cooling of a gas, i.e., throttling of a gas reduces its
temperature.
For µ < 0 ... throttling results in heating of a gas ... throttling of a gas increases its
temperature.
We have to find whether µ is greater than or less than zero.
Chemical Engineering Thermodynamics - I 5.99 Thermodynamic Properties of Pure Fluids

The Joule-Thomson coefficient of a gas is given by


∂T
µ =  
∂P H

The cyclic relation among H, T and P is


∂H ∂T  ∂P  = – 1
 ∂T  ∂P ∂H
     
P H T

∂H
 ∂P 
∂T = – 1  T 1 ∂y
∂P = – ... =  
 H    ∂P 
∂H ∂H  
∂x ∂xz
 ∂T  ∂H  ∂T  ∂y
 P  T  P  z

∂H
 ∂P 
∂T  T
∴ µ =   = –
∂PH ∂H
 ∂T 
  P

∂H
We have, CP =  
 ∂T  P

∂H
– 
 ∂P  T
∴ CPµ = ... (1)
We know that dH = T dS + V dP
Dividing by dP holding T constant, the above equation becomes
∂H ∂S
 ∂P  = T ∂P + V
   
T T

∂S ∂V
We know that ∂P = –  ... Maxwell’s relation
  T
 ∂T  P

∂H = – T ∂V + V
∴  ∂P   ∂T  ... (2)
 T  P
Combining Equations (1) and (2) gives
 ∂V  ∂V
– – T   + V T  –V
  ∂T P   ∂T P
µ = CP = CP ... (3)

∂V
Let us evaluate   .
 ∂T  P
The equation of state is
P (V – b) = RT
V = P T + b
R
 
Chemical Engineering Thermodynamics - I 5.100 Thermodynamic Properties of Pure Fluids

Differentiating this equation w.r.t. T at constant P, we get


∂V = R
 ∂T 
  P
P

∂V
Substituting for   in Equation (3) yields
 ∂T P
RT
P –V
µ = CP ... (4)
P (V – b) = RT
RT
∴ V–b = P
RT
Substituting for P by its value given above, Equation (4) becomes
V–b–V
µ = CP
b
∴ µ = –C ... Ans.
P
b is a positive constant, CP is always positive.
b
∴ C is always positive.
P
b
Since C is always positive, µ (given by – b/Cp) is negative, i.e., µ < 0.
P
Since µ is less than zero, throttling results in heating of the gas, i.e., the temperature of
the gas increases by throttling. ... Ans.
Example 5.35 :
Calculate the rise in pressure if liquid water is heated at constant volume from 303 K to
307.8 K. For water : β = 2 × 10–4 K–1 and κ = 4.8 × 10–4 MPa–1.
Solution :
1 ∂V
β = Coefficient of volume expansion = V   = 2 × 10–4 K–1
 ∂T  P
1 ∂V
κ = Coefficient of compressibility = – V   = 4.8 × 10–4 MPa–1
 ∂P  T
1 ∂V
V  ∂T 
β P
=
κ –   
1 ∂V
 V  ∂P  T

∂V
 ∂T 
 P β
∴ – = … (1)
∂V κ
 ∂P 
 T
Chemical Engineering Thermodynamics - I 5.101 Thermodynamic Properties of Pure Fluids

We have to find the rise in pressure, i.e., dP for the given rise in temperature, i.e., dT at
constant volume. So the partial derivatives in the above equation should be expressed in terms of
partial derivative of T w.r.t. P at constant volume.
For doing this, consider the relation f (P, V, T) = 0. Then
 ∂P  ∂V ∂T = – 1
∂V  ∂T  ∂P ... Cyclic relation
 T  P  V
∂V
 ∂P 
∂T = –1  T
∂P = –
 V  ∂P  ∂V ∂V
∂V  ∂T   ∂T 
 T  P  P
... Using reciprocal relation … (2)
Combining Equations (1) and (2) gives
∂V
 ∂T 
 P 1 ∂P
– = =  
∂V ∂T ∂TV
 ∂T  ∂P
 T  V

∂V
– 
 ∂T P ∂P β
∴ =   =
∂V ∂TV κ
 ∂T 
 T
∂P = β
∴ ∂T
 V κ
This can be written as
dP β
dT = κ
dT
dP = β
κ
2 × 10–4 (307.8 – 303)
dP = = 2 MPa
4.8 × 10–4
The rise in pressure is 2 MPa. ... Ans.
Example 5.36 :
Show that the Joule-Thomson coefficient is zero for ideal gases.
Solution :
The Joule-Thomson coefficient is given by
∂T
µ = µJT =  
∂P H
Chemical Engineering Thermodynamics - I 5.102 Thermodynamic Properties of Pure Fluids

We have to express µ in terms of the measurable quantities. To express the Joule-Thomson


coefficient of a gas in terms of the measurable quantities, consider the relation
f (H, T, P) = 0
Then
∂H ∂T  ∂P  = – 1
 ∂T  ∂P ∂H ... Cyclic relation
 P  H  T
∂H
 ∂P 
∂T 1  T
∴ µ =   = – = – … by reciprocal rule
∂PH ∂H  ∂P  ∂H
 ∂T  ∂H  ∂T 
 P  T  P

∂H
We have, CP =  
 ∂P  T

∂H
 ∂P 
 T
∴ µ = – C ... (1)
P

We know that, dH = T dS + V dP
Dividing by dP holding T constant, we get
∂H = T ∂S + V
 ∂P  ∂P
 T  T
∂S = – ∂V
We know that, ∂P  ∂T  ... Maxwell’s relation
 T  P
∂H = – T ∂V + V
Therefore,  ∂P   ∂T  ... (2)
 T  P
Combining Equations (1) and (2) gives
 ∂V   ∂V 
– – T   + V T  ∂T  – V
  ∂T  P  P 
µ = CP = CP ... (3)

For an ideal gas, PV = RT

V = P T
R

 
Differentiating w.r.t. T at constant P, we get
∂V = R
 ∂T  … (4)
 P P
Chemical Engineering Thermodynamics - I 5.103 Thermodynamic Properties of Pure Fluids

∂V
Substituting for   from Equation (4), Equation (3) becomes
 ∂T  P

RT – V
P 
µ = CP
RT
We have, PV = RT ∴ P = V
RT – V
P  V–V
∴ µ = CP = CP = 0 ... Ans.

Example 5.37 :
A rigid and insulated tank of 0.2 m3 volume is divided into two equal compartments by a
partition. One compartment initially contains 1 kmol CO2 at 500 K, while the second
compartment is evacuated. Determine the final temperature of CO2 in the tank if the
partition is removed. Assume CO2 to be a van der Waals gas with CV = 37 J/(mol·K) and
a = 363.077 × 10–3 Pa (m3/mol)2.
Solution :
Consider U = f (T, V)
∂U ∂U
dU =   dT +   dV
Then ... (1)
 ∂T  V
∂V T
We know that, dU = T dS – P dV
Dividing by dV holding T constant,
∂U = T  ∂S  – P
∂V ∂V
  T
  T

 ∂S  = ∂P ... Maxwell’s relation, the above equation becomes


With ∂V  
 T ∂TV
∂U = T ∂P – P
∂V ∂T
 T  V
∂U ∂U
Substituting for   and   , Equation (1) becomes
∂VT  ∂T V
 ∂P 
dU = CV dT + T   – P dV ... (2)
 ∂T  V
Given : CO2 obeys the relation
P + a  (V – b) = RT ... van der Waals equation of state
 V 2
P + V2 = V – b T
a R
 
Chemical Engineering Thermodynamics - I 5.104 Thermodynamic Properties of Pure Fluids

Differentiating this equation with respect to T at constant V, we get


∂P + 0 = R
∂T
  V
V–b

∂P R
∂T = V – b
 V
∂P
Substituting for   in Equation (2) gives
∂T V

dU = CV dT + V – b – P dV
RT
 
RT a
We have, V – b = P + V2
dU = CV dT + P + V2 – P dV
a

 
a
∴ dU = CV dT + V2 dV ... (3)
For this rigid and insulated tank, according to the first law of thermodynamics,
dU = U2 – U1 = Change in internal energy = 0 ... as dQ = 0
... insulated tank and dW = 0 ... free expansion
Integrating Equation (3), we get
U2 – U1 = CV (T2 – T1) – a V – V 
1 1
 2 1
With U2 = U1, i.e. U2 – U1 = 0 (Internal energy of CO2 is constant) we get
0 = CV (T2 – T1) – a V – V 
1 1
 2 1

CV (T2 – T1) = a V – V 
1 1
 2 1

T2 = T1 + C V – V  , here V is in m3 per mole


a 1 1
∴ ... (4)
V  2 1
State 1 : CO2 in one of the compartments of the tank.
V = Volume of tank = 0.2 m3 and the tank is divided into two equal
0.2
compartments, so V1 = 2 = 0.1 m3
Moles of CO2 = 1 kmol
0.1
∴ V1 = 1 = 0.1 m3/kmol = 10–4 m3/mol ... Specific volume
T1 = 500 K
State 2 : CO2 in the entire tank.
V = Volume of tank = 0.2 m3, T2 = Final temperature
Moles of CO2 = 1 kmol
∴ V2 = V = 0.2 m3 ... Final volume occupied by CO2
0.2
V2 = 1 = 0.2 m3/kmol = 2 × 10–4 m3/mol ... Specific volume
Chemical Engineering Thermodynamics - I 5.105 Thermodynamic Properties of Pure Fluids

We have, a = 363.077 × 10–3 Pa (m3/mol)2 ⇒ Pa·m6/mol2


CP = 37 J/(mol·K)
Substituting the numerical values of the terms involved in Equation (4), we get
363.077 × 10–3  1 1 
T2 = 500 + –
37 2 × 10–4 1 × 10–4
= 500 + 39.25
= 539.25 K
The final temperature of CO2 in the tank is 539.25 K ... Ans.
a 1 Pa·m6  1   1  ⇒ Pa·m ·K
3

CV V ... Units ⇒ mol m /mol J/(mol·K)


2 3 J
3
Pa·m ·K J·K
as J = N·m and Pa = N/m2 ... J = Pa·m3 ∴ J ⇒ J ⇒ K

Example 5.38 :
Show that the fugacity of a van der Waals gas is given by

ln f = V – b – RTV + ln V – b
b 2a RT
 
Estimate the fugacity of ethylene at 100 bar and 373 K.
Data : a = 0.453 J.m3/mol2, b = 0.574 × 10–4 m3/mol and V = 2.074 × 10–4 m3/mol at
100 bar and 373 K.
Solution :
For a van der Waals gas, we have
P + a  (V – b) = RT … equation of state for 1 mol.
 V 2
where a and b are van der Waals constants.
The above equation can be rewritten as
a RT
P + V2 = V – b
RT a
P = V – b – V2 … (1)
In this case, P is expressed as a function of V.
P
f 1
We have : ln o = RT ⌠
⌡ V dP … (2)
f
Po

It is known that f/P approaches unity at low pressure, i.e., fo/Po is virtually unity. Therefore,
ln (f/fo) may be replaced by ln (f/Po). Hence,
P
f 1
ln o = RT ⌠
⌡ V dP … (3)
P
Po
Chemical Engineering Thermodynamics - I 5.106 Thermodynamic Properties of Pure Fluids

To evaluate f requires a knowledge of ⌠


⌡ V dP which cannot be obtained from equations in
which P is expressed as a function of V. It is, therefore, advisable to rewrite ⌠
⌡ V dP in Equation
(3) as
⌠ V dP = ⌠
⌡ ⌡ d (PV) – ⌠
⌡ P dV
P V
∴ ⌠
⌡ V dP = PV – P V – ⌠
o o
⌡ P dV
Po Vo

where Vo is the molar volume of the gas corresponding to the low pressure Po. Since the
gas behaves ideally under this condition, we can replace PoVo in the above equation by
RT [PoVo = RT – ideal gas equation].
P V
⌠ P dV = PV – RT – ⌠
⌡ ⌡ P dV … (4)
Po Vo
RT a
We have, P = V – b – V2
Multiplying by V throughout, it becomes
RTV a
PV = V – b – V
Substracting RT from both sides, we get
RTV a
PV – RT = V – b – RT – V

= RT V – b – 1 – V
V a
 
= RT  V – b  – V
V – V + b a
 
RTb a
PV – RT = V – b – V … (5)
V
Substituting the value of P from Equation (1) in ⌠
⌡ P dV, we get
Vo
V V
⌠  RT a
⌡ P dV = ⌠
⌡ V – b – V2 dV
Vo Vo
V V
RT a
= ⌠
⌡ V – b dV – ⌠
⌡ V2 dV
Vo Vo
Integrating within the limits given, we get
V
⌠ V–b a a
⌡ P dV = RT ln Vo – b + V – Vo … (6)
Vo
Chemical Engineering Thermodynamics - I 5.107 Thermodynamic Properties of Pure Fluids

Since Vo is very large compared to b, Vo – b may be replaced by Vo [Vo – b ≈ Vo]. Further,


since Vo is the volume of a gas at a very low pressure Po at which the gas behaves ideally,
Vo may be replaced by RT/Po. Also, Vo being very large, a/Vo can be neglected. With these
simplifications, Equation (6) becomes
V
⌠ P dV = RT ln  o  + V
V–b a
⌡  V 
Vo
V
⌠ [V – b] Po + a
⌡ P dV = RT ln
 RT  V
… (7)
Vo

Substituting results from Equations (5) and (7) into Equation (4) gives
P
RTb a (V – b) Po a

⌡ P dV = V – b – V – RT ln RT – V
Po
P
RTb (V – b) Po 2a

⌡ P dV = V – b – RT ln RT – V … (8)
Po
P
⌠ P dV from (8), Equation (3) becomes
Substituting the value of ⌡
Po

f 1  RTb (V – b) Po 2a
ln o = RT V – b – RT ln – V
P  RT 
b 2a
ln f – ln Po = V – b – ln (V – b) – ln Po + ln RT – RTV

b RT 2a
∴ ln f = V – b + ln V – b – RTV … Ans.

Given : a = 0.453 J.m3/mol2, b = 0.571 × 10–4 m3/mol


V = 2.072 × 10–4 m3/mol
R = 8.31451 J/(mol.K) = 8.31451 × 10–5 (m3.bar)/mol.K
b 2a RT
ln f = V – b – RTV + ln V – b
0.571 × 10–4 2 × 0.453
= –
2.072 × 10 – 0.571 × 10
–4 –4 8.31451 × 373 × 2.072 × 10–4
8.31451 × 10–5 × 373
+ ln
2.072 × 10–4 – 0.571 × 10–4
= 0.3804 – 1.4099 + 5.33086 = 4.30136
∴ f = exp (4.30136) = 73.80 bar … Ans.
Note that for consistency of units, R in the term ln (RT/V – b) to be used is
8.31451 × 10–5 (m3.bar)/mol.K.
Chemical Engineering Thermodynamics - I 5.108 Thermodynamic Properties of Pure Fluids

Example 5.39 :
The proposed relationship between fugacity and pressure is f = P + aP2 where a is a
function of temperature only. Find an equation of state for the gas conforming to this relation.
Solution :
The relationship between fugacity and pressure is given as
f = P + aP2 … (1)
P
f Z–1
We have, ln P = ⌠
⌡ P dP … (2)
0

Introducing f from Equation (1) into Equation (2) gives


P
ln  P  = ⌠
P + aP2 (Z – 1)
  ⌡ P dP
0
P
dP
ln (1 + aP) = (Z – 1) ⌠
⌡ P
0

Taking derivatives of both sides


P
dP
d ln (1 + aP) = (Z – 1) d ⌠
⌡ P
0

a (Z – 1)
1 + aP = P
aP
1 + aP = Z – 1
aP
Z = 1 + 1 + aP

1 + aP + aP
Z = 1 + aP
1 + 2aP
Z = 1 + aP … Ans.
P
1 + 2aP f (Z – 1)
Extra : ⌠
Z = 1 + aP , we have : ln P = ⌡ P dP
0
P
f
ln P = ⌠ 1 + 2aP  dP
⌡  1 + aP – 1 P
0
P
f
ln P = ⌠ 1 + 2aP – 1 – aP dP
⌡  1 + aP  P
0
Chemical Engineering Thermodynamics - I 5.109 Thermodynamic Properties of Pure Fluids

P
f
ln P = ⌠  a 
⌡ 1 + aP dP
0
P
f
ln P = ⌠  a 
⌡ 1 + aP dP
0

dx
Let 1 + aP = x ∴ dP = a .

Limits : P = 0, x = 1 and P = P, 1 + aP = x.
1 + aP
f 1 a 1 + ap
∴ ln P = a ⌠
⌡ x dx = [ln x]1
1
f ...
ln P = ln (1 + aP) – ln 1 = ln (1 + aP), ln 1 = 0
f
ln P = ln (1 + aP)
∴ f/P = 1 + aP
f = P + aP2 … Given relationship between f and P.
Example 5.40 :
Derive an expression for the fugacity coefficient of a gas that obeys the following equation
of state :
RT a
P = V – b – T0.5 (V + b) V

Solution :
P
1
We have, ln f/Po ⌠ V dP
= RT ⌡
0
P V
and ⌠ V dP = PV – RT – ⌡
⌡ ⌠ P dV
Po Vo
V
RT ln f/Po = PV – RT – ⌠  RT – a 
⌡ V – b T 0.5 (V + b) V dV

Vo
V V
dV a dV
RT ln f/Po = PV – RT – RT ⌠
⌡ V – b – T0.5 ⌠
⌡ (V + b) V
Vo Vo
V V
V– b a 1 dV dV 
f/Po ⌠ –⌠
V – b T0.5 b ⌡ V ⌡ V + b
RT ln = PV – RT – RT ln o –
 Vo Vo 
a  V
– ln o 
V–b V+b
RT ln f/Po = PV – RT – RT ln o – ln
V – b T0.5 b  Vo V – b
Vo – b ≈ Vo = RT/Po, Vo – b ≈ Vo, with this the above equation becomes
Chemical Engineering Thermodynamics - I 5.110 Thermodynamic Properties of Pure Fluids

Po (V – b) a  V V o – b
RT ln f/Po = PV – RT – RT ln RT – 0.5 ln
T b  V o × V + b 
Po (V – b)
– T0.5 b ln V + b ,
a V
RT ln f/Po = PV – RT – RT ln RT  
since Vo – b = Vo
Po (V – b)
+ T0.5 b ln  V 
a V+b
RT ln f/Po = PV – RT – RT ln RT  
Po can be replaced by P

+ T0.5 b ln  V 
P (V – b) a V+b
RT ln f/P = PV – RT ln RT  
RT ln φ = PV – RT ln  RT  + T0.5 b ln  V 
P (V – b) a V+b
… Ans.
   
Example 5.41 :
Derive an expression to determine the fugacity of a gas obeying the van der Waals equation
of state.
Determine the fugacity of n-octane vapour at 427 K and 0.215 MPa.
Data : a = 3.789 Pa · (m3/mol)2, b = 2.37 × 10–4 m3/mol and V at 427 K and 0.215 MPa of
n-octane vapour = 15.675 × 10–3 m3/mol.
V–b a
ln f = Z – 1 – ln RT – RTV … obtain the expression in this form.

Solution :
For a van der Waals gas, the equation of state is
P + a  (V – b) = RT … for 1 mol
 V 2
where a and b are van der Waals constants.
The above equation can be written as (P as a function of V)
a RT
P + V2 = V – b
RT a
P = V – b – V2 … (1)
This equation expresses P as a function of V.
P
1
We have, ln f/fo = RT ⌠
⌡ V dP … (2)
Po

It is known that f/P approaches unity at low pressure, i.e., fo/Po is virtually unity. Therefore,
ln (f/fo) may be replaced by ln (f/Po). Hence, the above equation becomes
P
f 1
⌠ V dP
ln o = RT ⌡ … (3)
P
Po
Chemical Engineering Thermodynamics - I 5.111 Thermodynamic Properties of Pure Fluids

To evaluate f, we require a knowledge of ⌠


⌡ V dP which cannot be obtained from equations
in which P is expressed as a function of V. It is therefore advisable to rewrite ⌠
⌡ V dP in
Equation (3) as

⌡ V dP = ⌠
⌡ d (PV) – ⌠
⌡ P dV
P V
∴ ⌠
⌡ V dP = PV – PoVo – ⌠
⌡ P dV
Po Vo

where Vo is the volume of a gas corresponding to the low pressure Po. Since the
gas behaves ideally under this condition, we can replace PoVo in the above equation by
RT (PoVo = RT – for an ideal gas).
P V
∴ ⌠
⌡ P dV = PV – RT – ⌠
⌡ P dV … (4)
Po Vo
RT a
We have, P = V – b – V2
V
Let us evaluate ⌠
⌡ P dV
Vo
V V
⌠  RT a
⌡ P dV = ⌠
⌡ V – b – V2 dV
Vo Vo
V V
1 dV
= RT ⌠
⌡ V – b dV – a ⌠
⌡ V2
Vo Vo

Integrating within the limits given, it gives


P
V–b a a

⌡ P dV = RT ln Vo – b + V – Vo … (5)
Po

Since Vo is very large compared to b, Vo – b may be replaced by Vo [... Vo – b ≈ Vo].


Further, since Vo is the volume of a gas at a very low pressure Po at which the gas behaves
ideally, Vo may be replaced by RT/Po. Also Vo being very large, a/Vo can be neglected. With
these simplifications, Equation (5) becomes
V
V–b a

⌡ P dV = RT ln Vo + V
Vo
(V – b) Po a
= RT ln RT +V
Substituting this result in Equation (4), we get
P
(V – b) Po a
⌠ P dV = PV – RT – RT ln
⌡ RT –V … (6)
Po
Chemical Engineering Thermodynamics - I 5.112 Thermodynamic Properties of Pure Fluids

Combining Equations (3) and (6) gives


f 1  (V – b) Po a 
ln o = RT PV – RT – RT ln –V
P  RT 
PV (V – b) a
ln f – ln Po = RT – 1 – ln RT – ln Po – RTV
Since PV/RT = Z, the above equation becomes

ln f = Z – 1 – ln  RT  – RTV
V–b a
… Ans.
 
To evaluate the fugacity of n-octane vapour at 427 K and 0.215 MPa :
P = 0.215 MPa = 0.215 × 106 Pa
V = 15.675 × 10–3 m3/mol, a = 3.789 Pa · (m3/mol)2,
b = 2.37 × 10–4 m3/mol
and R = 8.31451 (m3.Pa)/mol.K
PV 0.215 × 106 × 15.675 × 10–3
Z = RT = = 0.9492
8.31451 × 427
15.675 × 10–3 – 2.37 × 10–4
ln f = 0.9492 – 1 – ln  
 8.31451 × 427 
3.789

8.31451 × 427 × 15.675 × 10–3
= 0.9492 – 1 – (– 12.3457) – 0.06808 = 12.37512
∴ f = exp (12.22682) = 204192.8 Pa = 0.2041928 MPa
≈ 0.2042 MPa … Ans.
b 2a RT
Also, we have ln f = V – b – RTV + ln V – b
2.37 × 10–4 2 × 3.789
= –
15.675 × 10–3 – 2.37 × 10–4 8.31451 × 427 × 15.675 × 10–3
8.31451 × 427
+ ln
15.675 × 10–3 – 2.37 × 10–4
= 0.01535 – 0.1362 + 12.3457
= 12.22485
∴ f = exp (12.22485) = 203790.95 = 203791 Pa
= 203791 × 10–6 MPa ≈ 0.2038 MPa … Ans.
This answer is almost equal to the previously obtained.
Example 5.42 :
Derive an expression for the fugacity coefficient of a gas obeying the following equation of
state :
Z = a + bP + cP2
where a, b and c are empirical constants and P is in bar.
Determine the fugacity of oxygen at 20°C (293 K) and 100 bar.
Data : a = 1.0, b = – 0.753 × 10–3 and c = 0.15 × 10–5.
Chemical Engineering Thermodynamics - I 5.113 Thermodynamic Properties of Pure Fluids

Solution :
The fugacity coefficient (f/P) in terms of the compressibility factor is given by
P
f
ln P = ⌠ Z – 1
⌡  P  dP … (1)
0

We have, Z = a + bP + cP2
Substituting for Z, Equation (1) becomes
P
a + bP + cP – 1 dP
f 2
ln P = ⌠
⌡ P 
0

Integrating it gives,
P P P
f (a – 1)
ln P = ⌠
⌡ P dP + ⌠⌡ b dP + ⌠
⌡ cP dP
0 0 0

cP2
= (a – 1) ln P + bP + 2

f cP2
ln P = ln φ = (a – 1) ln P + bP + 2 … Ans.

Let us find the fugacity of oxygen at 100 bar.


P = 100 bar, a = 1.0, b = – 0.753 × 10–3 and c = 0.15 × 10–5
f cP2
ln P = (a – 1) ln P + bP + 2

= (1 – 1) ln 100 – 0.753 × 10–3 × 100 + 0.15 × 10–5 × (100)2/2


= – 0.0678
ln f – ln P = – 0.0678
ln f – ln 100 = – 0.0678
ln f – 4.605 = – 0.0678
ln f = 4.5372, f = e4.5372 ∴ f = 93.43 bar
The fugacity of oxygen gas at 100 bar is 93.43 bar … Ans.
Example 5.43 :
Derive an expression for the fugacity coefficient of a gas that obeys the equation of state
PV B C
RT = 1 + V + V2 , where P is in bar.
Evaluate the fugacity of argon at 0°C (273 K) and 1 bar using the expression obtained,
given that
B = – 21.13 × 10–6 m3/mol and
C = 1054 × 10–12 m6/(mol)2
Chemical Engineering Thermodynamics - I 5.114 Thermodynamic Properties of Pure Fluids

Solution :
The equation of state provided is
PV B C
RT = 1 + V + V2 … (1)

This equation is called the Leiden form or pressure explicit form of the virial equation.
The virial equation of state mentioned above is expressed as a power series in reciprocal
volumes.
The virial equation of state can also be written as a power series in pressure as
PV
RT = 1 + B'P + C'P
2 … (2)

This equation is called the Berlin form or volume explicit form of the virial equation.
The virial coefficients B' and C' are related to B and C as given below :
B C – B2
B' = RT and C' = (RT)2

Substituting for B' and C', Equation (2) becomes


PV BP (C – B2) P2
RT = 1 + RT + (RT)2 … (3)

PV
But we have, Z = RT … compressibility factor.

Therefore, Equation (3) can be written as


BP (C – B2) P2
Z = 1 + RT + (RT)2 … (4)

The fugacity coefficient (f/P) is related to the compressibility factor through the following
equation :
P
f
ln P = ⌠ Z – 1
⌡  P  dP … (5)
0

Substituting Z from Equation (4) in Equation (5), we get


P
 BP (C – B ) P 
f 2 2
ln P = ⌠
⌡ 1 + RT + (RT)2 – 1 dP/P
0
P P
f B dP (C – B2) P
ln P = ⌠
⌡ RT + ⌠
⌡ (RT)2 dP
0 0

Integrating gives,
f BP (C – B2) P2
ln P = RT + 2 (RT)2 … Ans.
Let us now evaluate the fugacity of argon using the above equation.
P = 1 bar
B = – 21.13 × 10–6 m3/mol
Chemical Engineering Thermodynamics - I 5.115 Thermodynamic Properties of Pure Fluids

C = 1054 × 10–12 m6/(mol)2


T = 273 K and R = 8.31451 m3.kPa/(kmol.K)
R = 8.31451 (m3.Pa)/mol.K
We have, 105 Pa = 1 bar
(m3.Pa) 1 bar
∴ R = 8.31451 mol.K × 105 Pa = 8.31451 × 10–5 (m3.bar)/mol.K

f BP (C – B2) P2
ln P = RT + 2 (RT)2

BP (C – B2) P2
ln f – ln P = RT + 2 (RT)2

(– 21.13 × 10–6) × 1 [1054 × 10–12 – (– 21.13 × 10–6)2] × (1)2


ln f – ln 1 = +
8.31451 × 10 × 273
–5 2 (8.31451 × 10–5)2 (273)2
ln f – 0 = – 9.309 × 10–4 + 1.456 × 10–6
ln f = 0 – 9.309 × 10–4 + 1.456 × 10–6 = – 9.2944 × 10–4
f = exp (– 9.2944 × 10–4) = 0.99907 ≈ 0.9991 bar
The fugacity of argon gas at 273 K and 1 bar is 0.9991 bar … Ans.
Example 5.44 :
Calculate the fugacity of gaseous ammonia if its density at 200°C (473 K) and 50 bar is
24.3 kg/m3.
Solution :
Let us first calculate the molar volume of ammonia.
ρ of ammonia at 473 K and 50 bar = 24.3 kg/m3.
Mol. Wt. of ammonia = MNH3 = 17 kg/kmol

The molar volume of ammonia at 473 K and 50 bar is


V = MNH3 / ρ
17 kg/kmol
= 24.3 kg/m3 = 0.69958 ≈ 0.70 m3/kmol

The fugacity of ammonia under the given conditions is given by


VP2
f = RT

where, V = molar volume = 0.70 m3/kmol


P = 50 bar = 50 × 105 N/m2 = 50 × 102 kN/m2 = 50 × 102 kPa
T = 473 K
R = 8.31451 m3.kPa / (kmol.K)
Chemical Engineering Thermodynamics - I 5.116 Thermodynamic Properties of Pure Fluids

0.70 × (50 × 102)2


f =
8.31451 × 473
= 4449.79 kPa ≈ 4449.79 kN/m2
= 4449.79 × 103 N/m2
1 bar
= 4449.79 × 103 N/m2 × 105 N/m2

f = 44.4979 ≈ 44.50 bar … Ans.


Example 5.45 :
Carbon tetrachloride boils at 76.75°C (349.75 K) at 1 bar. The latent heat of vaporization of
CCl4 is 194.8 kJ/kg. What would be its boiling point at 2 bar ?
Solution :
The integrated Clausius-Clapeyron equation is
s
P2 ∆Hv  1 1 
ln s = R T1 – T2
P1
s
where, P1 = 1 bar = 105 N/m2 = 102 kN/m2 = 102 kPa
s
P2 = 2 bar = 2 × 102 kPa
∆Hv or ∆Hvap = latent heat of vaporisation = 194.8 kJ/kg
Mol. Wt. of CCl4 = 154 kg/kmol
∆Hv = 194.8 kJ/kg × 154 kg/kmol = 29999.2 kJ/kmol
T1 = 349.75 K, T2 = ?
R = 8.31451 kJ/(kmol.K)
Substituting the values, we get
2 × 102 = 29999.2  1 – 1 
ln  
1 × 102 8.31451 349.75 T2

0.693 = 3608.054 2.859 × 10–3 – T 


1
 2

1
1.9207 × 10–4 = 2.859 × 10–3 – T
2

1
T2 = 2.859 × 10 – 1.9207 × 10
–3 –4

∴ T2 = 374.96 K
The boiling point of CCl4 is 101.96 °C [374.96 K] … Ans.
Chemical Engineering Thermodynamics - I 5.117 Thermodynamic Properties of Pure Fluids

Example 5.46 :
The vapour pressure of liquid chlorine at 0 °C (273 K) is 3.71 bar and the molar volume
under these conditions is 6.01 × 10–3 m3/mol. Calculate the fugacity of liquid chlorine at 273 K.
Solution :
The fugacity of the liquid is equal to that of the vapour (with which it is in equilibrium)
at 273 K and 3.71 bar with V equal to 6.01 × 10–3 m3/mol.
Therefore, the fugacity of liquid chlorine can be calculated using the following equation :
VP2
f = RT
where, P = 3.71 bar, V = molar volume = 6.01 × 10–3 m3/mol
T = 273 K
m3.Pa m3.Pa 1 bar
R = 8.31451 mol.K = 8.31451 mol.K × 105 Pa
= 8.31451 × 10–5 m3.bar/(mol.K)
6.01 × 10–3 × (3.71)2
f = = 3.644 bar
8.31451 × 10–5 × 273
The fugacity of liquid chlorine is 3.644 bar at 273 K, which is somewhat less than its vapour
pressure. … Ans.
Example 5.47 :
The experimental P-V data for benzene at 402°C (675 K) from very low pressures upto about
75 bar, may be represented by the equation V = 0.0561 (1/P – 0.0046), where V is the molar
volume in m3/mol and P is in bar. Find the fugacity of benzene at 1 bar and 675 K.
Solution :
The molar volume of benzene at 1 bar is
V = 0.0561 [1/P – 0.0046]
= 0.0561 [1/1 – 0.0046]
= 0.05584 m3/mol
Given : T = 675 K, P = 1 bar
m3.Pa m3.Pa 1 bar
T = 8.31451 mol.K = 8.31451 mol.K × 105 Pa
= 8.31451 × 10–5 (m3.bar)/mol.K
The fugacity of benzene at 1 bar and 675 K is
P2V
f = RT
(1)2 × 0.05584
=
8.31451 × 10–5 × 675
= 0.9949 ≈ 0.995 bar … Ans.
Example 5.48 :
Utilise the following data for 0 °C (273 K) to calculate the fugacity of carbon monoxide at
100 and 400 bar pressure.
P 25 50 100 200 400 800
PV/RT 0.9890 0.9792 0.9741 1.0196 1.2482 1.8057
Chemical Engineering Thermodynamics - I 5.118 Thermodynamic Properties of Pure Fluids

Solution :
P
ln f/P = ⌠
⌡ (α/RT) dP … (1)
0

RT
α = V– P

α V 1 PV – 1 1
= – = … (2)
RT RT P RT  P
We are provided with PV/RT at various P values. So evaluate α/RT using Equation (2) at the
various P values.
P, bar 25 50 100 200 400 800
PV/RT 0.9890 0.9792 0.9741 1.0196 1.2482 1.8057
α/RT – 4.4 × 10–4 – 4.16 × 10–4 –2.59 × 10–4 0.98 × 10–4 6.205 × 10–4 10.07 × 10–4
The value of the integral on the RHS of Equation (1) is the area under the curve of a plot of
α/RT v/s P.
Construct a plot of α/RT v/s P.
(1) To find the fugacity of carbon monoxide at 100 bar : Refer to Fig. E 5.48.
Area under the curve between 0 and 100 bar pressure = Area A-B-C-D = – 389 mm2
f
ln P = area bounded by A-B-C-D × (scale of y-axis) × (scale of x-axis)
1 × 10–4 100
ln f – ln P = – 389 mm2 × 10 mm × 10 mm

ln f – ln (100) = – 0.0389
ln f – 4.6052 = – 0.0389
ln f = 4.6052 – 0.0389 = 4.5663
f = exp. (4.5663) = 96.187 ≈ 96.20 bar … Ans.
(2) To find the fugacity of carbon monoxide at 400 bar : Refer Fig. E 5.47.
Area under the curve between 0 and 400 bar = Area A-B-C-E + Area E-F-G
= – 462 + 705 = 243 mm2
400
f
ln P = ⌠
α
⌡ RT dP
0
ln f – ln P = Area between 0 and 400 bar × (Scale of Y-axis) × Scale of X-axis)
1 × 10–4 100
ln f – ln 400 = 243 mm2 × 10 mm × 10 mm
ln f – 5.9915 = 0.0243
ln f = 6.0158
∴ f = exp. (0.0158) ( i.e.‚ e6.0158) = 409.85 ≈ 409.9 bar
The fugacity of carbon monoxide at 400 bar and at 273 K is 409.9 bar … Ans.
Chemical Engineering Thermodynamics - I 5.119 Thermodynamic Properties of Pure Fluids

-4
11´10
-4
10´10
-4
9´10
-4
8´10
-4
7´10
-4
6´10 F

-4
5´10
-4
4´10
2
Area E-F-G = 705 mm (+ve)
a / RT

-4
3´10
-4
2´10
-4
1´10
100 200 400 600 800
0,0
A D E G
P
-4
-1´10
2
-4 Area A-B-C-E = 462 mm (-ve)
-2´10
-4 C
-3´10 2
Area A-B-C-D = 389 mm (-ve)
-4
-4´10
-4 B
-5´10

Fig. E 5.48
Chemical Engineering Thermodynamics - I 5.120 Thermodynamic Properties of Pure Fluids

Example 5.49 :
Using the following data, calculate the fugacity of nitrogen gas at 100 and 800 bar and at
0°C (273 K).
P, bar 50 100 200 400 800 1000
PV/RT 0.9846 0.9846 1.0365 1.2557 1.7959 2.0641
Solution :
P, bar 50 100 200 400 800 1000
PV/RT 0.9846 0.9846 1.0365 1.2557 1.7959 2.0641
α –3.08 × 10–4 – 1.54 × 10–4 1.825 × 10–4 6.3925 × 10–4 9.949 × 10–4 1.064 × 10–3
RT
(10.64 × 10–4)
P
f α
ln P = ⌠
⌡ RT dP … (1)
0
RT
α = V – P … residual volume
α V 1 PV  1
RT = RT – P = RT – 1 P … (2)
Based on the PV/RT values at various pressures, evaluate α/RT using Equation (2).
To evaluate the integral on the RHS of Equation (1) (to find f) between zero and any pressure P,
we have to plot α/RT against P.
(A) To find the fugacity of nitrogen gas at 100 bar and 273 K : Refer to Fig. E 5.49.
Area under the curve (A-B-C) between 0 and 10 bar pressure = – 304 mm2
(–ve sign as the curve is below the X-axis)
100
f α
ln P = ⌠ ⌡ RT dP
0
= Area under the curve (A-B-C) × (Scale of Y-axis) × (Scale of X-axis)
1 × 104 100
ln f – ln P = – 304 mm2 × 10 mm × 10 mm = – 0.0304
ln f – ln (100) = – 0.0304
ln f – 4.6052 = – 0.0304 ∴ ln f = 4.6052 – 0.0304 = 4.5748
f = exp. (4.5748) is 97 bar
The fugacity of nitrogen at 100 bar is 97 bar … Ans.
(B) To find the fugacity at 800 bar and at 273 K : Refer to Fig. E 5.49.
The value of the integral of α/RT between 0 and 800 bar pressure is equal to the area under
the curve between 0 and 800 bar pressure on both sides of the X-axis.
Area under the curve
between 0 and 800 bar pressure = Area under A-B-D + Area under DEF
= – (333) + (4350) = 4017 mm2
800
f
ln P = ⌠
⌡ (α/RT) dP
0
= Area × (Scale of Y-axis) × (Scale of X-axis)
1 × 10–4 100
= 4017 mm2 × 10 mm × 10 mm
Chemical Engineering Thermodynamics - I 5.121 Thermodynamic Properties of Pure Fluids

Fig. E 5.49
Chemical Engineering Thermodynamics - I 5.122 Thermodynamic Properties of Pure Fluids

ln f – ln P = 0.4015
ln f – ln (800) = 0.4015
ln f – 6.6846 = 0.4015
ln f = 7.0861
f = exp (7.0861) = 1195.23 ≈ 1195 bar
The fugacity of nitrogen at 800 bar and 273 K is 1195 bar. … Ans.
[Please note : a small error in plotting and measuring the area under the curve may result in a
large error in the answer. So precision is required.]
Example 5.50 :
From the following compressibility data for oxygen at 200 K, determine the fugacity of
oxygen at 100 bar and 200 K.
P, bar 1 4 7 10 40 70 100
Z 0.99701 0.98796 0.9788 0.96956 0.8734 0.7764 0.6871
Solution :
P
ln f/P = ⌠ Z – 1
We have, ⌡  P  dP
0

The integral in the above equation is found out graphically by plotting (Z – 1)/P against P.
At P = 1, Z = 0.99701
Z–1 0.99701 – 1
∴ P = 1 = – 0.00299 = – 2.99 × 10–3

Similarly, evaluate (Z – 1)/P at other values of P and tabulate.


P, bar 1 4 7 10 40 70 100

Z–1
( )
P
– 2.99 × 10–3 – 3.01 × 10–3 – 3.0286 × 10–3 – 3.044 × 10–3 – 2.816 × 10–3 – 3.19 × 10–3 – 3.129 × 10–3

Construct a plot of (Z – 1)/P v/s P.


Refer Fig. E 5.50.
Area under the curve between P = 0 and P = 100 bar is 9090 mm2.
ln f/P = Area × (Scale of X-axis) × (Scale of Y-axis)
10
= 9090 mm2 × 10 mm ×
–1.0 × 10–3
 30 mm 
= – 0.303
ln f – ln P = – 0.303
ln f – ln (100) = – 0.303
ln f – 4.60517 = – 0.303, ln f = 4.29857
∴ f = exp (4.29857) = 73.594 ≈ 73.6 bar … Ans.
Chemical Engineering Thermodynamics - I 5.123 Thermodynamic Properties of Pure Fluids

P
0,0 10 20 30 40 50 60 70 80 100
(z-1/P)

-4
-1´10 2
Area = 9090 mm

-4
-2´10

-4
-3´10

-4
-4´10

Fig. E 5.50
Chemical Engineering Thermodynamics - I 5.124 Thermodynamic Properties of Pure Fluids

Example 5.51 :
The experimental pressure-volume data for benzene at 675 K from very low pressures upto
about 75 atm, may be approximated by the equation

V = 54.405 P – 0.0046
1
 
where P is in atmospheres and V is in cubic meters per kilogram-mole.
(A) What is the fugacity of benzene at 1 atm and 675 K ?
Take R = 0.0806 m3.atm/(kmol.K).
(B) Assuming the fugacity of benzene to be unity at 1 atm and 675 K, calculate its fugacity
at 75 atm and 675 K.
Solution :
We have, RT d ln f = V dP

V = 54.405 P – 0.0046 , m3/mol


1
Given :
 
At P = 1 atm, we have to find f.
T = 675 K and R = 0.0806 m3.atm/(kmol.K)

RT d ln f = 54.405 P – 0.0046 dP
1
 

0.0806 × 675 d ln f = 54.405 P – 0.0046 dP


1
 

54.405 d ln f = 54.405 P – 0.0046 dP


1
 

d ln f = P – 0.0046 dP
1
 
dP
d ln f = P – 0.0046 dP

dP
d ln f = d ln P – 0.0046 dP, since P = d ln P

d ln f – d ln P = – 0.0046 dP
∴ d ln (f/P) = – 0.0046 dP
Integrating gives,
ln f2/P2 – ln f1/P1 = – 0.0046 (P2 – P1)
We know that when P approaches zero, then f/P approaches 1.0.
∴ As P1 → 0, f1/P1 = 1.0
Chemical Engineering Thermodynamics - I 5.125 Thermodynamic Properties of Pure Fluids

Therefore,
ln (f2/P2) – ln (1) = – 0.0046 (P2 – 0)
ln f2/P2 – 0 = – 0.0046 P2
Given : P = P2 = 1 atm
∴ ln f2/1 = ln f2 = – 0.0046 × 1 = –0.0046
ln f2 = – 0.0046
f2 = exp (– 0.0046) = 0.9954 atm
The fugacity of benzene at 1 atm and 675 K is 0.9954 atm. … Ans. (A)

V = 54.405 P – 0.0046
1
OR :
 
Given : P = 1 atm
V = 54.405 1 – 0.0046 = 54.1547 m3/kmol
1

 
The fugacity of benzene can be calculated by using the relation
VP2
f = RT
54.1547 × (1)2
=
0.0806 × 675
= 0.995399 ≈ 0.9954 atm … Ans. (A)
To find the fugacity at 75 atm and 675 K :
We have :
ln f2/P2 – ln f1/P1 = – 0.0046 (P2 – P1)
Given : At P = P1 = 1 atm, f = f1 = 1
ln f2/P2 – ln 1/1 = – 0.0046 (P2 – 1)
ln f2 – ln P2 = – 0.0046 (P2 – 1)
Given : P2 = 75 atm
ln f2 – ln 75 = – 0.0046 (75 – 1)
ln f2 – 4.3175 = – 0.3404
ln f2 = 3.9771
∴ f2 = exp (3.9771) = 53.36 atm … Ans. (B)
Example 5.52 :
Steam boilers are provided with pressure relief valves which do not permit the pressure in
the boiler to build-up beyond a certain level. It is required to design a boiler to supply saturation
steam at 473 K and so it is necessary to design a pressure relief valve that releases the pressure
if it exceeds the saturation pressure at 473 K. Find the saturation pressure at 473 K. The normal
boiling point of water is 373 K and its latent heat of vaporisation at 373 K is 2257 kJ/kg.
Chemical Engineering Thermodynamics - I 5.126 Thermodynamic Properties of Pure Fluids

Solution :
The normal boiling point is the boiling point at 101.325 kPa (1 atm).
s
∴ T1 = 373 K and P1 = 101.325 kPa
s
P1 – Saturation pressure/vapour pressure of water at 373 K
∆Hv or ∆Hvap = latent heat of vaporisation of water
= 2257 kJ/kg
= 2257 kJ/kg × 18 kg/kmol … Mol. Wt. of H2O = 18 kg/kmol
= 40626 kJ/kmol
s
P2 = ? - saturation pressure at T2 = 473 K.
The Clausius-Clapeyron equation is
Ps2 ∆Hv  1 1 
ln  s = R T – T , R = 8.31451 kJ/(kmol.K)
P1  1 2

ln P2 – ln (101.325) = 8.31451 373 – 473


s 40626 1 1
 
s
ln P2 – 4.6183 = 4886.16 [2.68096 × 10–3 – 2.11416 × 10–3]
s
ln P2 = 4.6183 + 2.76947 = 7.38777
s
∴ P2 = exp (7.38777) = 1616 kPa = 1.616 MPa
The saturation pressure of water at 473 K is 1.616 MPa. … Ans.
Example 5.53 :
The pressure inside a pressure cooker is 200 kPa. Find the boiling point of water inside the
cooker by making use of the data provided.
Data : The normal boiling point of water is 373 K and its latent heat of vaporisation is
2257 kJ/kg at 373 K.
Solution :
Given : ∆Hv = latent heat of vaporisation of water
= 2257 kJ/kg
= 2257 kJ/kg × 18 kg/kmol … Mol. Wt. of H2O = 18 kg/kmol
= 40626 kJ/kmol
T1 = 373 K – the normal boiling point of water, i.e., the boiling point at
P = 1 atm = 101.325 kPa
s
∴ P1 = 101.325 kPa
s
P2 = 200 kPa, T2 = ?
R = 8.31451 kJ/(kmol.K)
Chemical Engineering Thermodynamics - I 5.127 Thermodynamic Properties of Pure Fluids

The integrated Clausius-Clapeyron equation is


s
P2 ∆Hv 1 1
ln  s = R T – T 
P1  1 2
Substituting the values,

ln 101.325 = 8.31451 373 – T 


200 40626 1 1
   2

0.67998 × 8.31451 1
40626 = 2.6809 × 10–3 – T
2

1
1.3916 × 10–4 = 2.6809 × 10–3 – T
2

1
T2 = 2.6809 × 10 – 1.3916 × 10 = 2.54174 × 10
–3 –4 –3

∴ T2 = 393.43 K
The boiling point of water inside the cooker is 393.43 K. … Ans.
Example 5.54 :
The vapour pressure data of benzene are as given below :
T (K) 280.6 288.4 299.1 315.2 333.6 353.1
Ps (bar) × 102 5.33 8.00 13.30 26.70 53.33 100
Making use of the above data,
(i) Estimate the latent heat of vaporisation.
(ii) The vapour pressure of benzene at 393 K.
Solution :
(i) 1 bar = 105 Pa = 102 kPa
s 102 kPa
∴ P1 = 5.33 × 10–2 bar = 5.33 × 10–2 bar × 1 bar = 5.33 kPa

T, K 280.6 288.4 299.1 315.2 333.6 353.1


Ps , kPa 5.33 8.0 13.30 26.70 53.33 100
1/T 3.56 × 10–3 3.467 × 10–3 3.343 × 10–3 3.1726 × 10–3 2.9976 × 10–3 2.832 × 10–3
ln Ps 1.673 2.08 2.59 3.285 3.976 4.605
The Clausius-Clapeyron equation is
∆Hv 1
ln Ps = – R T + constant
 
where, ∆Hv = latent heat of vaporisation, kJ/kmol
Ps = vapour pressure in kPa
R = 8.31451 kJ/(kmol.K)
Chemical Engineering Thermodynamics - I 5.128 Thermodynamic Properties of Pure Fluids

It follows from the above equation that a plot of the logarithm (natural) of the vapour
pressure, i.e., ln Ps against the reciprocal of the absolute temperature, i.e., 1/T, should yield a
straight line of slope – ∆Hv/R

• Convert Ps to ln Ps and T to 1/T and tabulate.


• Plot ln Ps v/s 1/T, which gives slope = – ∆Hv/R and from the slope, evaluate the mean
latent heat of vaporisation in the given temperature range.

Refer to Fig. E 5.54.

Slope = – 4034.48
∴ – 4034.48 = – ∆Hv/R = – ∆Hv/8.31451

∴ ∆Hv = 4034.48 × 8.31451 = 33544.7 ≈ 33545 kJ/kmol

The latent heat of vaporisation of benzene in the given temperature range is 33545 kJ/kmol,
i.e., 33.545 × 103 kJ/kmol. … Ans. (i)
(ii) To find the vapour pressure of benzene at 393 K :
The integrated form of Clausius-Clapeyron equation is
Ps2 ∆Hv  1 1 
ln  s = R T – T
P1  1 2

where, ∆Hv = 33545 kJ/kmol


s
P1 = 100 kPa, T1 = 353.1 K … from the data provided
s
P2 = ?, T2 = 393 K, R = 8.31451 kJ/(kmol.K)

Ps2
ln  s = 8.31451 353.1 – 393
33545 1 1
P1  
s s
ln (P2 ) – ln (P1 ) = 4034.51 [2.832 × 10–3 – 2.544 × 10–3]
s
ln (P2 ) – ln (100) = 1.1619
s
ln (P2 ) – 4.605 = 1.1619
s
ln (P2 ) = 4.605 + 1.1619 = 5.7669
s
∴ P2 = exp (5.7669) = 319.54 kPa

The vapour pressure of benzene at 393 K is 319.54 kPa … Ans. (ii)


Chemical Engineering Thermodynamics - I 5.129 Thermodynamic Properties of Pure Fluids

5.0

4.0

- 58.5 ´ (1/20)
Slope =
s
In P

-3
29 ´ (1´10 /40)
3.0 = - 4034.48

2.0

1.0

0,0 -3 -3 -3 -3
1´10 2´10 3´10 4´10
1/T

Fig. E 5.54
Chemical Engineering Thermodynamics - I 5.130 Thermodynamic Properties of Pure Fluids

Example 5.55 :
Prove the following :
∂Cp ∂Cp
  – Cp  
∂µ
  = –µ
 ∂P T  ∂T P ∂TP
where µ is the Joule-Thomson coefficient.
∂Cp
Evaluate the derivative   at P = 1 bar and T = 33 K.
 ∂P T
Given : Cp = 6.557 + 6.1477 × 10–2 T – 2.148 × 10–7 T2, kJ/(kmol.K)
µ = – 0.1975 + 138/T – 319 P/T2, K/bar
Solution :
∂S ∂S Cp
We have : Cp = T   ∴   = T
∂TP ∂TP
∂S = Cp
∂T
 P T
Differentiating w.r.t. P at constant T gives
∂2S ∂2S 1 ∂Cp
= = T   … (1)
∂P ∂T ∂T ∂P  ∂P T
The Maxwell's relation is
∂S ∂V
∂P = –  ∂T 
 T  P
Differentiating w.r.t. T gives
∂S ∂2V
= –  2 … (2)
∂T ∂P  ∂T P
Combining Equations (1) and (2), we get
1 ∂Cp ∂2V
 
T  ∂P T = –  ∂T2 
 P
∂Cp ∂2V
  = –T  ∂T2  … (3)
 ∂P T  P
T (∂V/∂T)P – V
We have, µ = µJT = Cp
∴ µCp = T (∂V/∂T)P – V … (4)
Differentiating Equation (4) w.r.t. T at constant P gives
∂Cp ∂2V ∂V ∂V
 + Cp   = T  2  + 1 ×   –  
∂µ
µ
 ∂T P ∂TP  ∂T P  ∂T P  ∂T P
∂Cp
 + Cp   = T  2 
∂µ ∂2V
∴ µ … (5)
 ∂T P ∂TP  ∂T P
Chemical Engineering Thermodynamics - I 5.131 Thermodynamic Properties of Pure Fluids

Combining Equations (3) and (5), we get


∂Cp ∂Cp
 + Cp   = –  
∂µ
µ
 ∂T P ∂TP  ∂P T
∂Cp  ∂Cp ∂µ 
∴   = – µ   + Cp   
 ∂P T   ∂T P ∂TP
∂Cp ∂Cp
 – Cp  
∂µ
  = –µ … Ans.
 ∂P T  ∂T P ∂TP
To evaluate the derivative (∂Cp/∂P)T
138 319 P
Given : µ = – 0.1975 + T – T2 , K/bar
Differentiating this equation w.r.t. T at constant P gives
∂µ = 0 + 318 – 1  – 319 P – 2  = – 318 + 638
∂T  T 2  T 3 … (6)
 P T2 T3
Cp = 6.577 + 6.1477 × 10–2 T – 2.148 × 10–7 T2
∂Cp
  = 0 + 6.1477 × 10–2 × 1 – 2.148 × 10–7 (2T)
 ∂T P
= 6.1477 × 10–2 – 4.296 × 10–7 T … (7)
Substituting (∂µ/∂T)P from Equation (6) and (∂Cp/∂T)P from Equation (7), Equation for
(∂Cp/∂P)T becomes
∂Cp
  = – – 0.1975 + T – T2  [6.1477 × 10–2 – 4.296 × 10–7 T]
138 319 P
 ∂P T

– [6.577 + 6.1477 × 10–2 T – 2.148 × 10–7 T2] – T2 + T3 


318 638
 
Substituting T = 333 K and P = 1 bar, we get
∂Cp
  = – – 0.1975 + 333 – (333)2  [6.1477 × 10–2 – 4.296 × 10–7 × (333)]
138 319 × 1
 T
∂P  

– [6.577 + 6.1477 × 10–2 (333) – 2.148 × 10–7 (333)2] – (333)2 + (333)3


138 638
 
= – [0.2140] [0.06133] – [27.033] [– 1.227 × 10–3]
= – 0.01312 + 0.03317 = 0.020 (kJ/kmol) (bar)–1 … Ans.
Example 5.56 :
Calculate the change in internal energy when 10 kmol of CO2 gas expands isothermally from
a pressure of 101325 kPa to 101.325 kPa at 373 K. The corresponding molar volumes are
0.215 m3/kmol and 30.53 m3/kmol respectively. Assume that CO2 obeys the following relation :
P + 365 (V – 0.043) = RT
 V2 
Chemical Engineering Thermodynamics - I 5.132 Thermodynamic Properties of Pure Fluids

Solution : Given : n = moles of CO2 = 10 kmol


V1 = initial molar volume = 0.215 m3/kmol
V2 = final molar volume = 30.53 m3/kmol
Isothermal expansion at T = 373 K.
The change in internal energy in an isothermal process is given by
∂U = T ∂P – P … effect of V on U at constant T
∂V ∂T … (1)
 T  V
Find (∂P/∂T)V from the given equation of state.
CO2 gas obeys the relation
P + 365 (V – 0.043) = RT
 V2 
Rearranging gives
RT 365
P = V – 0.043 – V2
Differentiating this equation with respect to T at constant V, we get
∂P R (1) R
∂T = V – 0.043 + RT (0) – 0 = V – 0.043 … (2)
 V
Substituting for (∂P/∂T)V from Equation (2) in Equation (1) gives
∂U = RT
∂V –P
 T V – 0.043
Substituting for P, this equation becomes
∂U = RT RT 365 365
∂V – + = V2
 T V – 0.043 V – 0.043 V2
∂U 365
Therefore, = V2
∂V
365
dU = V2 dV
U2 V2 30.53
dV dV
∴ ⌠
⌡ dU = 365 ⌠
⌡ V2 = 365 ⌠
⌡ V2
U1 V1 0.215

V–2+1 30.53
∆U = U2 – U1 = 365 –2 + 1
 0.215
1 30.53
∆U = – 365 V  1 1 
 0.215 = – 365 30.53 – 0.215
= 1685.7 kJ/kmol
The change in internal energy for 10 kmol CO2 gas is
∆U = 10 × 1685.7 = 16857 kJ … Ans.
Chemical Engineering Thermodynamics - I 5.133 Thermodynamic Properties of Pure Fluids

Example 5.57 :
Carbon tetrachloride boils at 349.75 K at 1 bar. What would be the vapour pressure of
carbon tetrachloride at 374.97 K bar ?
The latent heat of vaporisation of CCl4 is 154.8 kJ/kg.
Solution :
The Clausius-Clapeyron equation, in the integrated form, (assuming ∆Hv to be constant in
the temperature range provided) is
s
P2 ∆Hv  1 1 
ln s = R T1 – T2
P1
s
where, P1 = saturation pressure = 1 bar
s
P2 = saturation pressure
s
T1 = boiling point at P1 = 349.75 K
s
T2 = boiling point at P2 = 374.97 K
∆Hv = latent heat of vaporisation = 194.8 kJ/kg
Mol. Wt. of CCl4 = 154 kg/kmol
∆Hv = 194.8 kJ/kg × 154 kg/kmol = 29999.2 kJ/kmol
R = 8.31451 kJ/(kmol.K)
Substituting the values of the various parameters, we get

ln P2 – ln P1 = 8.31451 349.75 – 374.97


s s 29999.2 1 1
 
s
ln P2 – ln 1 = 0.6927
s
ln P2 – 0 = 0.6927
s
ln P2 = 0.6927
s
P2 = Exp. (0.6927)
= 1.9996 bar ≈ 2 bar
The boiling point of CCl4 at 374.97 K is 2 bar. … Ans.
Example 5.58 :
The vapour pressure data of acetone are as given below :
T (K) 290.1 320.5 350.9
Ps (kPa) 21.525 77.445 201.571
Calculate the latent heat of vaporisation of acetone by making use of the data provided.
Chemical Engineering Thermodynamics - I 5.134 Thermodynamic Properties of Pure Fluids

Solution :
T (K) 290.1 320.5 350.9
Ps (kPa) 21.525 77.445 201.571
The integrated Clausius-Clapeyron equation is
Ps2 ∆Hv  1 1 
ln  s = R T – T
P1  1 2
where ∆Hv is the latent heat of vaporisation.
s
(1) At T1 = 290.1 K, P1 = 21.525 kPa
s
At T2 = 320.5 K, P2 = 77.445 kPa
R = 8.31451 kJ/(kmol.K)
∆Hv
ln 21.525 = 8.31451 290.1 – 320.5
77.445 1 1
   
∆Hv
1.2803 = 8.31451 [3.4471 × 10–3 – 3.120 × 10–3]
∴ ∆Hv = 32543.769 ≈ 32543.8 kJ/kmol
s
(2) At T1 = 320.5 K, P1 = 77.445 kPa
s
At T2 = 350.9 K, P2 = 201.571 kPa
∆Hv
ln  77.445  = 8.31451 320.5 – 350.9
201.571 1 1
   
∆Hv
0.9566 = 8.31451 [3.120 × 10–3 – 2.8498 × 10–3]
∴ ∆Hv = 29436.19 ≈ 29436.2 kJ/kmol
s
(3) At T1 = 290.1 K, P1 = 21.525 kPa
s
At T2 = 350.9 K, P2 = 201.571 kPa
∆Hv
ln  21.525  = 8.31451 290.1 – 350.9
201.571 1 1
   
∆Hv
2.2369 = 8.31451 [3.4471 × 10–3 – 2.8498 × 10–3]
∴ ∆Hv = 31138 kJ/kmol
32543.8 + 29436.2 + 31138
Average ∆Hv = 3
= 31039.33 ≈ 31039 kJ/kmol
The latent heat of vaporisation of acetone is 31039 kJ/kmol … Ans.
The graphical method can also be used to find the latent heat of vaporisation.
Chemical Engineering Thermodynamics - I 5.135 Thermodynamic Properties of Pure Fluids

Example 5.59 :
Prove the following Maxwell's relations :
 ∂T  = – ∂P
(i) ∂V ∂S
 S  V
∂T = ∂V
(ii) ∂P  ∂S 
 S  P
∂V = – ∂S and
(iii)  ∂T  ∂P
 P  T
 ∂S  = ∂P
(iv) ∂V  
 T ∂TV
Solution :
(i) For a pure substance undergoing an infinitesimal reversible process/change,
we have
dU = T dS – P dV … (1)
At constant volume, dV = 0. Therefore, the above equation becomes
dU = T dS
∂U = T
or  ∂S 
 V
Differentiating the above equation with respect to volume (V) at constant S, we get
∂2U  ∂T 
=   … (2)
∂S ∂V ∂VS
At constant entropy, dS = 0. Therefore, Equation (1) becomes
dU = – P dV
∂U = – P
or ∂V
 S
Differentiating this equation with respect to S at constant V gives
∂2U ∂P
= –  … (3)
∂V ∂S ∂SV
Since the order of differentiation of the second mixed partial derivative [∂2U/∂V ∂S] is
immaterial, we can write Equation (3) as
∂2U ∂2U ∂P
= = –  … (4)
∂V ∂S ∂S ∂V ∂SV
Comparing Equations (2) and (4), we get
∂T = – ∂P
∂V  ∂S  … Ans. (i)
 S  V
Chemical Engineering Thermodynamics - I 5.136 Thermodynamic Properties of Pure Fluids

(ii) For a pure substance undergoing an infinitesimal reversible change, we have


dH = dU + P dV + V dP, since H = U + PV
i.e., dH = T dS + V dP … differential form of H … (1)
[dU = dQ – dW ⇒ dU = dQ – P dV ⇒ dU + P dV = dQ ⇒
dQ = T dS Therefore, dU + P dV = T dS]
At constant pressure, i.e., when dP = 0, the above equation becomes
dH = T dS
∂H = T
or  ∂S 
 P
Differentiating this equation with respect to P at constant S gives
∂2H  ∂ ∂H  ∂T
=   ∂S   = ∂P … (2)
∂P ∂S   PS  S
∂P
At constant S, i.e., when dS = 0, Equation (1) becomes
dH = V dP
∂H = V
or  ∂P 
 S
Differentiating this equation with respect to S at constant P, we get
∂2H  ∂ ∂H  ∂V
=    =   … (3)
∂S ∂P   SP  ∂S P
∂S ∂P
Since the order of differentiation for the second mixed partial derivative (∂2H/∂P ∂S) is
immaterial, we can write Equation (3) as
∂2H ∂2H ∂V
= =   … (4)
∂S ∂P ∂P ∂S  ∂S P
Comparing Equations (2) and (4), we get
∂T = ∂V
 ∂P    … Ans. (ii)
 S  ∂S P
(iii) For a pure substance undergoing an infinitesimal reversible change/process,
we have
dG = dH – T dS – S dT
i.e., dG = V dP – S dT … differential of G … (1)
At constant temperature, i.e., when dT = 0, Equation (1) becomes
dG = V dP
∂G = V
or  ∂P 
 T
Chemical Engineering Thermodynamics - I 5.137 Thermodynamic Properties of Pure Fluids

Differentiating this equation with respect to T at constant P gives


∂2G  ∂ ∂G  ∂V
=    =   … (2)
∂T ∂P ∂T  ∂P TP  ∂T P
At constant pressure, i.e., when dP = 0, Equation (1) becomes
dG = – S dT
∂G = – S
or  ∂T 
 P
Differentiating this equation with respect to P at constant T yields
∂2G  ∂ ∂G  ∂S
=     = –  … (3)
∂P ∂T ∂P  ∂T PT ∂PT
Since the order of differentiation for the second mixed partial derivative (∂2G/∂T ∂P) is
immaterial, we can write
∂2G ∂2G ∂S
= = –  … (4)
∂P ∂T ∂T ∂P ∂PT
Comparing Equations (2) and (4), we get
∂V = – ∂S
 ∂T  ∂P … Ans. (iii)
 P  T
(iv) For a pure substance undergoing an infinitesimal reversible process or change, we have
dA = dU – T dS – S dT
i.e., dA = – S dT – P dV … differential form of A … (1)
[dU = dQ – P dV ⇒ dU = T dS – P dV ⇒ dU – T dS = – P dV]
At constant volume, i.e., when dV = 0, Equation (1) becomes
dA = – S dT
∂A = – S
or  ∂T 
 V
Differentiating this equation with respect to V at constant T gives
∂2A  ∂ ∂A   ∂S 
=     = –  … (2)
∂V ∂T ∂V  ∂T VT ∂VT
At constant temperature, i.e., when dT = 0, Equation (1) becomes
dA = – P dV
∂A = – P
or ∂V
 T
Differentiating this equation with respect to T at constant V gives
∂2V  ∂ ∂A  ∂P
=     = –  … (3)
∂T ∂V   TV
∂T ∂V ∂TV
Chemical Engineering Thermodynamics - I 5.138 Thermodynamic Properties of Pure Fluids

 ∂2A  is
Since the order of differentiation for the second mixed partial derivative  
∂V ∂T
immaterial, we can write
∂2A ∂2A ∂P
= = –  … (4)
∂T ∂V ∂V ∂T ∂TV

Comparing Equations (2) and (4), we get


 ∂S  ∂P
–  = – 
∂VT ∂TV
 ∂S  =  ∂P 
i.e., ∂V ∂T … Ans. (iv)
 T  V
Extra : Writing the Maxwell's relations using the Mnemonic or Born Diagram :
1 V T
1'

3' 2' P
2 S 3

Fig. 5 (A)

[When the arrows in the square are pointing in the same direction while looking to the square
from any side, there is no need to add a negative sign in the Maxwell's relation. If it is not so
(i.e., if the arrows in the square are not pointing in the same direction), then we have to add a
negative sign in the Maxwell's relation.]

Refer to Fig. 5 (A). Here we are looking from the bottom (upward) and the arrows are
pointing upward, i.e., in the same direction. So, no need to add a negative sign in the Maxwell's
relation.

Derivative of 1 w.r.t. 2 at constant 3 = Derivative of 1' w.r.t. 2' at constant 3'.


∂V = ∂T
 ∂S  ∂P
 P  S

(B) Refer to Fig. 5 (B). We are looking from the top (downward), the arrows in the square
are pointing in the same direction (upward). So no need to add a negative sign in the Maxwell's
relation.
Chemical Engineering Thermodynamics - I 5.139 Thermodynamic Properties of Pure Fluids

2 V ' T 3
3 2'

1' P
1 S

Fig. 5 (B)

Derivative of 1 w.r.t. 2 at constant 3 = Derivative of 1' w.r.t. 2' at constant 3'

 ∂S  = ∂P
∂V  
 T ∂TV

(C) Refer to Fig. 5 (C). Here we are looking from the LHS of square. While looking from
the LHS (towards the square) one arrow (the arrow pointing towards T) is pointing upwards and
the other arrow in the square (i.e., the arrowing pointing towards V) is pointing downwards. That
is, the arrows in the square are pointing in the opposite directions. Hence, in this case it is
necessary to add a negative sign in the Maxwell's relation.

2 V T 1
3'
Do
wn
p
U

2' 1'
3 S P

Fig. 5 (C)

(Derivative of 1 w.r.t. 2 at constant 3) = – (Derivative of 1' w.r.t. 2' at constant 3').

 ∂T  = – ∂P
∂V ∂S
 S  V

(D) Refer to Fig. 5 (D). While looking from the RHS of the square, one arrow in the square,
i.e., the arrow pointing towards V is pointing upwards, while the arrow pointing towards T is
pointing downwards. Since the arrows in the square are not pointing in the same direction,
we have to add a negative sign in the Maxwell's relation.
Chemical Engineering Thermodynamics - I 5.140 Thermodynamic Properties of Pure Fluids

1 V T 2
3'

Up
wn
Do

1' 2'
S P 3

Fig. 5 (D)
(Derivative of 1 w.r.t. 2 at constant 3) = – (Derivative of 1' w.r.t. 2' at constant 3')
∂V = – ∂S
 ∂T  ∂P
 P  T
❐❐❐
Chapter ... 6

REFRIGERATION
REFRIGERATION
• Refrigeration is the process of lowering and maintaining the temperature of a system
below that of the surroundings.
• Refrigeration is the process of cooling of a system to a temperature below that of the
surroundings / surrounding atmosphere. This requires continuous absorption or
removal of heat from the system at a low temperature which is achieved by evaporation
of a liquid in a steady-state flow process. The vapour thus formed is converted into its
liquid state for its re-evaporation more commonly by compression followed by
condensation (or by absorption in a liquid of low volatility).
• Refrigeration implies the production and maintenance of low temperatures by
continuously absorbing heat at a low temperature level and rejecting it at a high
temperature level (i.e., it implies the continuous removal of heat from a system whose
temperature is always below the temperature of the surroundings). Since heat cannot
flow spontaneously from a body at low temperature to a body at a high temperature,
external work is required to be done on the system to achieve refrigeration.
A refrigerator / refrigerating system is a device which maintains a system at a temperature
lower than the temperature of the surroundings (by operating in a cycle).
A system which is maintained at a temperature lower than that of its surroundings is called a
refrigerated system.
APPLICATIONS OF REFRIGERATION
1. Manufacture of ice.
2. Manufacture and storage of ice creams.
3. Comfort air conditioning of houses, hotels, buses, railways, offices and hospitals.
4. Industrial air conditioning.
5. Processing of food products and beverages.
6. Freezing, storage and transportation of food products and beverages.
7. Liquefaction of process gases and dehydration of gases.
8. Lubricating oil purification.
(6.1)
Chemical Engineering Thermodynamics - I 6.2 Refrigeration

9. Manufacture of synthetic rubbers, textiles, dyes, plastics, dimethyl terephthalate,


chlorine and caprolactum.
10. Preservation of foods, fruits, fish, meat, drinks, medicine, milk, etc. for domestic
purpose.
11. To remove heat of chemical reactions.
TYPES OF REFRIGERATORS / REFRIGERATING MACHINES / SYSTEMS
Refrigerators are classified as :
(i) Air refrigerators, and
(ii) Vapour refrigerators.
In the air refrigerating systems, the working fluid is air which is non-condensible. Due to the
non-condensibility of air, it does not undergo change of state and thus only sensible heat changes
take place in these systems.
In the vapour refrigerating systems, the work fluid is vapour which is condensible and
evaporative. Due to the condensibility and evaporativeness of the vapour used, it undergoes
changes of state from liquid to vapour and back to liquid and thus latent heat changes take place
in these systems.
Vapour refrigerators are further classified as :
(i) Vapour compression refrigerating machines, and
(ii) Vapour absorption refrigerating machines.
The vapour compression systems are electric power operated units, whereas the vapour
absorption systems are heat operated units. These two systems utilize different means for
compressing the refrigerant. In the vapour compression systems, a compressor is used, whereas
in the vapour absorption systems, an absorber - generator assembly is used.
REFRIGERATING EFFECT AND REFRIGERATION UNIT CAPACITY
The amount of heat extracted from the space to be refrigerated in a given time is termed as
the refrigerating or refrigeration effect.
The capacity of a refrigeration unit or refrigerator is rated in tons (or tonnes) of refrigeration.
• A ton of refrigeration (or one ton of refrigeration) is the refrigeration produced
(or refrigerating effect produced) by melting one ton of ice at a temperature of 0°C in one
day (in 24 hours).
• A ton of refrigeration is defined as the amount of heat removed from one ton of pure water at
0°C to form ice at 0°C in one day (in 24 hours).
• It is the rate of heat removal / absorption equivalent to the removal of 12000 Btu or 12660 kJ
per hour (in SI units), i.e., a ton of refrigeration or one ton of refrigeration is equivalent to
heat removal at a rate of 12660 kJ/h.
Chemical Engineering Thermodynamics - I 6.3 Refrigeration

1 ton of refrigeration corresponds to a refrigeration rate of 12660 kJ/h.


The enthalpy of fusion of one ton of ice at 32°F (0°C) is roughly 288000 Btu (British thermal
unit).

1 ton of refrigeration = (288000 Btu) ×


1055 × 10–3 kJ ×  1 
 1 Btu  24 h
= 12660 kJ/h
REVERSED CARNOT CYCLE
A device that operates on the reversed Carnot cycle is called a Carnot refrigerator or a heat
pump. Fig. 6.1 shows a reversed Carnot cycle, which is used for producing refrigeration.
P-V diagram and T-S diagram for this cycle are shown in Fig. 6.1 (b).
Q1

4 3
Condenser

Expansion
WE Compressor WC
Engine

1 Evaporator 2

Q2

(a) Components of a reversed Carnot cycle (Carnot refrigerator)

T1 4 3
4

3
P T
T2 2
1
1
2
5 6
V S
(b) P-V and T-S diagrams for the reversed Carnot cycle
Fig. 6.1
Chemical Engineering Thermodynamics - I 6.4 Refrigeration

In the reversed Carnot cycle, the following processes are carried out sequentially.
In the reversed cycle, the refrigerant (working fluid) is first compressed reversibly and
adiabatically, i.e., isentropically in process 2-3. In the process 3-4, heat is removed from the
working fluid at constant temperature T1 and the working fluid then expands isentropically. The
temperature of the working fluid falls from T1 to T2 and work output of WE is obtained. Finally,
the working fluid absorbs heat from the evaporator / refrigerator reversibly by evaporation at
constant temperature T2 and the cycle is repeated.
Process 1-2 : Isothermal heat absorption
Process 2-3 : Isentropic compression
Process 3-4 : Isothermal heat rejection
Process 4-1 : Isentropic expansion
During the isothermal heat absorption, an amount of heat Q2 is absorbed at the lower
temperature T2. It is represented by the area 1-2-6-5 on the T-S diagram.
Q2 = T2 (S2 – S1) ... (6.1)
During the isothermal heat rejection, an amount of heat Q1 is rejected at the higher
temperature T1. It is represented by the area 3-4-5-6 on the T-S diagram.
Q1 = T1 (S3 – S4) ... (6.2)
The net work done is represented by the enclosed area 1-2-3-4.
Wnet = WC – WE = Q1 – Q2 = (T1 – T2) (S2 – S1) ... (6.3)
The coefficient of performance of a Carnot refrigerator is given by
Refrigerating effect
COP = Net work required
Heat absorbed at the lower temperature
= Net work required
Q2 Q2 T2 (S2 – S1)
= W = Q – Q = (T – T ) (S – S )
net 1 2 1 2 2 1

T2
COP = T – T ... (6.4)
1 1

where T1 is the temperature of heat rejection and T2 is the temperature of heat absorption
(T1 > T2).
Equation (6.4) gives the maximum possible value of the COP for any refrigerator operating
between given values of T1 and T2. It shows clearly that the refrigerating effect per unit of work
required (i.e., the COP) decreases as T2 (the temperature of the refrigerator) decreases and as
T1 (the temperature of heat rejection) increases.
For the same T1 or T2, the COP increases with the decrease in the temperature difference
(T1 – T2). i.e., the higher COP is obtained if the temperatures T1 and T2 are closer to each other.
REFRIGERATION SYSTEMS
(i) Air refrigeration system,
(ii) Vapour compression refrigeration system,
(iii) Vapour absorption refrigeration system.
Chemical Engineering Thermodynamics - I 6.5 Refrigeration

AIR REFRIGERATION SYSTEM (BELL - COLEMAN CYCLE)


In the Bell-Coleman cycle, heat absorption and heat rejection are taking place at constant
pressures instead of at constant temperatures as in the Carnot cycle.
In an air refrigerating system, air is used as the refrigerant to absorb / remove heat from the
cold space (the space to be refrigerated) and to reject the same into the surrounding atmosphere
which is at a higher temperature than the cold space. The working fluid, air, does not undergo
phase change in the system.
The components of an air refrigeration system are : compressor, cooler (heat exchanger),
expander and refrigerator. Fig. 6.2 (a) shows a flow diagram of the air refrigeration system. The
working of air-refrigeration cycle is represented on P-V and T-S diagrams as shown in
Fig. 6.2 (b).
Q1

4 3
Cooler
P2

WE
Expander Compressor WC

P1

Refrigerator

1 2
Q2

(a) Air refrigeration system

(b) P-V and T-S diagrams


Fig. 6.2
Chemical Engineering Thermodynamics - I 6.6 Refrigeration

In this system, air from the refrigerator / cold space is compressed isentropically (reversibly
and adiabatically) to a pressure P2 and temperature T3. The hot compressed air delivered by the
compressor is cooled in the cooler using water at constant pressure P2. Heat removed from the air
causes a great decrease in its volume. The high pressure cooled air is expanded in the expander
isentropically. In the expander, the work is produced at the expense of the enthalpy of air which
results in a large drop in the air temperature. Due to isentropic expansion, the air is cooled to a
temperature below that of the cold space. The low temperature air enters the cold space /
refrigerator and absorbs heat at constant pressure P1. The air leaving the cold space again enters
the compressor and the cycle is repeated.
Process 1-2 : Heat absorption from cold space at constant pressure (P1)
Process 2-3 : Isentropic compression of air
Process 3-4 : Cooling of air / heat rejection at constant pressure (P2)
Process 4-1 : Isentropic expansion of air
T3 - Temperature of compressed air after compressor
T1 - Temperature of air entering the refrigerator
T4 - Temperature of cooled air leaving the cooler
T2 - Temperature of air entering the compressor
Heat absorbed from the cold space :
· ·
Q2 = m Cp (T2 – T1) , kJ/s
or Q2 = Cp (T2 – T1) , kJ/kg ... i.e., per kg of air
Heat rejected to the surroundings / removed from the air in the cooler :
Q1 = Cp (T3 – T4) , kJ/kg ... i.e., per kg of air
· ·
or Q1 = m Cp (T2 – T1) , kJ/s
The net work required (i.e., done on air) is given by
Wnet = WC – WE = Q1 – Q2
= Cp (T3 – T4) – Cp (T2 – T1), in kJ/kg ... (6.5)
Wnet may be expressed in kJ/s.
The coefficient of performance of an air refrigeration system is given by
Q2 Cp (T2 – T1)
COP = W = C (T – T ) – C (T – T )
net p 3 4 p 2 1

Refrigerating effect
= Net work required
(T2 – T1)
COP = (T – T ) – (T – T ) ... (6.6)
3 4 2 1

Air undergoes isentropic compression from 2 to 3.


Chemical Engineering Thermodynamics - I 6.7 Refrigeration

For an isentropic compression in which the pressure changes from P1 to P2 , we have


γ–1
T3 P2 γ
= P  ... (6.7)
T2  1
This equation relates the initial and final temperatures of air, undergoing an isentropic
compression, to initial and final pressures.
Similarly, for an isentropic expansion, we can write
γ–1
γ
T4 = P2
T  P  ... (6.8)
 1  1
From Equations (6.7) and (6.8), we get
γ–1
P2 γ
T3 = T2 P  ... (6.9)
 1
γ–1
P2 γ
and T4 = T1 P  ... (6.10)
 1
Subtracting Equation (6.10) from Equation (6.9), we get
γ–1
γ
T3 – T4 = (T2 – T1) P 
P2
... (6.11)
 1
Substituting Equation (6.11) into Equation (6.6), we get
(T2 – T1)
COP = γ–1
P2 γ
(T2 – T1) P  – (T2 – T1)
 1
1
COP = γ–1 ... (6.12)
γ
P2
P  –1
 1
Substituting Equation (6.8) in Equation (6.12) gives
1
COP = T
4
T1 – 1
T1
COP = T – T ... (6.13)
4 1

Or substituting Equation (6.7) in Equation (6.12), we get


1 T2
COP = T = T –T ... (6.14)
3 3 2
T2 – 1
Chemical Engineering Thermodynamics - I 6.8 Refrigeration

The above equations of COP are valid only when the compression and expansion processes
are isentropic.
If the compression and expansion follow a polytropic process [PVn = C] , then
n
WC = n – 1 (P3V3 – P2V2) ... (6.15)

n
and WE = n – 1 (P4V4 – P1V1) ... (6.16)

The net work required by the air refrigeration system is


Wnet = WC – WE
n n
Wnet = n – 1 (P3V3 – P2V2) – n – 1 (P4V4 – P1V1) ... (6.17)

For one mole of a gas : PV = RT


∴ P3V3 = RT3, P2V2 = RT2 and so on
With this, the net work required becomes
n nR
Wnet = n – 1 R (T3 – T2) – n – 1 (T4 – T1)

nR
i.e., Wnet = n – 1 (T3 – T2 + T1 – T4), kJ/kmol ... (6.18)

where R = 8.31451 kJ/(kmol·K)


The molecular weight of air is 28.84 ≈ 29 kg/kmol.
With this Equation (6.18) becomes
nR
Wnet = 29 (n – 1) (T3 – T2 + T1 – T4), kJ/kg ... (6.19)

n
WC = n – 1 (P3V3 – P2V2) ... (6.20)

For one mole of an ideal gas, PV = RT, P3V3 = RT3 and P2V2 = RT2
nR
∴ WC = n – 1 (T3 – T1), kJ/kmol ... (6.21)
nR
or WC = 29 (n – 1) (T3 – T1), kJ/kg
nR
Similarly, WE = n – 1 (T4 – T1), kJ/kmol
nR
or WE = 29 (n – 1) (T4 – T1), kJ/kg ... (6.22)
The refrigerating effect is
Q2 = Cp (T2 – T1), kJ/kg, if Cp is in kJ/(kg·K) ... (6.23)
or Q2 = Cp (T2 – T1), kJ/kmol, if Cp is in kJ/(kmol·K) ... (6.24)
Chemical Engineering Thermodynamics - I 6.9 Refrigeration

We have, Cp – Cv = R
Cp R
Cv – 1 = Cv
Cp
But, Cv = γ
R R
∴ γ–1 = C = C –R
v p
(γ – 1) (Cp – R) = R
Cp (γ – 1) – γR + R = R

Cp = 
γ 
∴  R ... (6.25)
γ – 1
With this, the equation for Q2 becomes

Q2 = 
γ 
 R (T2 – T1) , kJ/kmol ... (6.26)
 1
γ –
where R is 8.31451 kJ/(kmol·K)
Q2 = 29 
1 γ 
or  R (T2 – T1) , kJ/kg ... (6.27)
 1
γ –
The COP of the system is
Q2
COP = W
net
Substituting Q2 from Equation (6.26) and Wnet from Equation (6.18), we get

Wnet = 
γ  R (T2 – T1)
 · nR
γ – 1
(n – 1) (T3 – T2 + T1 – T4)
(T2 – T1)
i.e., Wnet = ... (6.28)
γ 
 γ  n – 1 (T3 – T2 + T1 – T4)
– 1 n
 
The advantages of the air refrigeration systems are :
(i) The working fluid, air, is available at free cost.
(ii) There is no danger of fire or toxic effects in case of air leakage.
(iii) The weight of air refrigeration system per ton of refrigeration is less compared to that
of other refrigeration systems. Because of this, the air refrigeration system is used in
air crafts.
The disadvantages of the air refrigeration systems are :
(i) The COP of air refrigeration system is low in comparison with that of other
refrigeration systems. Hence, the running cost of the system is high.
(ii) The quantity of air to be circulated in this system is more compared with the
refrigerants used in other refrigeration systems (because of heat transfer in the form of
sensible heat).
Chemical Engineering Thermodynamics - I 6.10 Refrigeration

VAPOUR COMPRESSION REFRIGERATION SYSTEM


The modern refrigerating plants work on the vapour compression system.
Fig. 6.3 (a) shows the simplified diagram of a vapour compression refrigeration plant /
system.
The four basic components in the system are the compressor, condenser, expansion /
throttling valve and evaporator. The processes that are completed in one cycle are : compression,
condensation, expansion and vaporization.
Fig. 6.3 (b) shows the vapour compression cycle on a T-S diagram and Fig. 6.3 (c) shows the
same cycle on a P-H diagram.
Q1

Condenser
4 3

Expansion Compressor W or WC
valve

Evaporator
1 2

Q2
Space to be cooled
or refrigerated

(a) Vapour compression system

Q1
P1 4 3
3 3'
4 P1
Temperature

T1
P or In P

3'
P2 1 2
P2 Q2
T2 2
1

Entropy (S) Enthalpy (H)

(b) Vapour compression cycle (c) Vapour compression cycle


on T-S diagram on P-H diagram
Fig. 6.3
Chemical Engineering Thermodynamics - I 6.11 Refrigeration

Process 1-2 : Reversible constant pressure evaporation / vaporization - heat absorption


Process 2-3 : Reversible adiabatic compression, i.e., isentropic compression of the vapour
Process 3-4 : Reversible constant pressure condensation - heat rejection
Process 4-1 : Irreversible adiabatic throttling process (expansion process), during which
enthalpy remains constant (H4 = H1)
State of refrigerant at 2 : Saturated vapour at 3 : Superheated vapour at 3' : Saturated vapour
at 4 : Saturated liquid and at 1 : A mixture of liquid and vapour, i.e., a two-phase mixture.
The low pressure refrigerant vapour is compressed in the compressor. In compressing the
vapour, heat of compression is added to the vapour as the pressure is increased. The superheated
vapour leaving the compressor enters the condenser, in which heat is removed from the
refrigerant causing it to condense to a liquid at the pressure held in the condenser. In the
condenser, latent heat of condensation is removed from the refrigerant vapour with the help of a
cooling medium such as cooling water. The liquid refrigerant flows from the condenser to the
expansion valve in which a Joule-Thomson expansion occurs, both P and T of the liquid
refrigerant are reduced. This expansion results in evaporation of some liquid refrigerant and
cooling of both liquid and vapour refrigerant to the temperature of the evaporator. A mixture of
liquid and vapour refrigerant from the expansion valve enters the evaporator. In the evaporator,
heat is absorbed from the space to be cooled or refrigerated (or the fluid to be cooled or
refrigerated). Due to the absorption of heat, the liquid portion of the refrigerant is converted to
low pressure saturated vapour. From the evaporator, the saturated vapour refrigerant enters the
compressor. The cycle is thus completed.
Refer to the vapour compression cycle presented on the T-S and P-H diagrams. The
assumptions made for drawing these diagrams are : (i) the vapour leaving the evaporator is dry
and saturated, (ii) the compression process is isentropic (reversible and adiabatic) and (iii) no
subcooling of the refrigerant in the condenser.
The vapour leaving the evaporator is dry and saturated represented by point 2. It is then
isentropically compressed till it becomes superheated (point 3). The isentropic compression is
represented by the line 2-3. In the condenser, vapour is first cooled to remove superheat
i.e., vapour is first desuperheated represented by the curved line 3-3' and then condensed at
constant T and P to get saturated liquid (4). The process of cooling and condensation is
represented by the line 3-4. In the condenser, heat Q1 is transferred to the cooling water.
The saturated liquid at 4 undergoes an adiabatic throttling process. For this, enthalpy remains
constant (the line 4-1 in Fig. 6.3 represents the same). During this expansion process, both P and
T are reduced to those in the evaporator. Throttling results in the partial vaporization of the
liquid refrigerant, so we obtain a mixture of liquid and vapour refrigerant, the point 1
representing the mixture lies in the two phase region. The liquid-vapour mixture at 1 enters the
evaporator where it absorbs / extracts heat (Q2) from the space to be cooled or refrigerated at
constant pressure.
Chemical Engineering Thermodynamics - I 6.12 Refrigeration

The expansion valve throttles the refrigerant to a pressure, the saturation temperature
corresponding this pressure is below the temperature of the space to be cooled. So, heat flow,
because of temperature difference, from the space (or refrigerator), which gets cooled or
refrigerated, to the refrigerant. The evaporator thus produces cooling or refrigerating effect by
absorbing heat (Q2) from the space to be refrigerated or refrigerator.
Due to absorption of heat, the liquid portion of refrigerant gets vaporized / evaporated and
we get low pressure saturated vapour which then enters the compressor to complete the cycle.
The vaporization / evaporation process occurring at constant P is represented by the line 1-2.
The refrigerant absorbs heat from the cold space (cold body), which acts as a source of latent
heat and thus it gets converted to vapour from liquid. It rejects heat to the cooling water
(hot body) in the condenser and gets converted from vapour to liquid, i.e., gets condensed.
Therefore, the vapour compression refrigeration system is a latent heat pump since it pumps
latent heat from the cold space and delivers it to the cooling water (cooling medium).
[The process 4-1 is an irreversible adiabatic expansion of the vapour (through the throttle
valve) because of which both P and T of the liquid refrigerant are decreased. States 4 and 1 are
equilibrium states, but the process 4 to 1 is irreversible. Therefore, actually it cannot be shown
on property diagrams or should be shown by a dotted line. But still it is shown by a continuous
line for the sake of understanding.]
ANALYSIS OF VAPOUR COMPRESSION SYSTEM
Refer to Fig. 6.3 (b) and (c). The following assumptions are made in drawing the T-S and
P-H diagrams of a vapour compression cycle.
(i) The refrigerant coming out the refrigerator is dry and saturated.
(ii) There is no subcooling of the refrigerant in the condenser.
(iii) The compression is isentropic, i.e., without friction and heat transfer.
(iv) There are no pressure losses (i.e., pressure drop) due to fluid friction in the system.
(v) The first law for a steady flow process may be written for each component in the cycle,
neglecting KE and PE changes.
(vi) Isentropic compression results in a superheated vapour (represented by point 3).
(vii) There is no heat exchange between the system and the surrounding except in the
evaporator and condenser.
1. Compressor :
Input energy = Output energy
H2 + WC = H3
∴ WC = H3 – H2 , kJ/kg
Chemical Engineering Thermodynamics - I 6.13 Refrigeration

2. Condenser :
H3 = Q1 + H4
Q1 = H3 – H4 , kJ/kg
3. Evaporator :
H1 + Q2 = H2
Q2 = H2 – H1 , kJ/kg
H1, H2, H3 and H4 are the specific enthalpies of refrigerant at the respective positions /
locations as shown in Fig. 6.3 (c).
The refrigerating effect or the amount of heat removed or absorbed from the cold space per
kg of refrigerant is given by
Q2 = H2 – H1 , kJ/kg ... (6.29)
Work required for the refrigerating effect per kg of refrigerant is given by
WC = W = H3 – H2 , kJ/kg ... (6.30)
The coefficient of performance of this vapour compression cycle is given by
Refrigerating effect Q2
COP = Work required = W
H2 – H1 H2 – H4
= H – H = H – H , since H1 = H4 ... (6.31)
3 2 3 2

·
Let m be the rate of circulation of refrigerant (i.e., the mass flow rate of refrigerant) in kg/s.
The heat rejected in the condenser to the cooling water (i.e., heat removed from the
refrigerant in the condenser by the cooling water) is given by
Q1 = H3 – H4 = H3 – H1 , since H4 = H1 , kJ/kg ... (6.32)
The rate of heat removal in the condenser is given by
·
Q1 = m (H3 – H4), kJ/s
·
If mw is the rate of water circulation in the condenser, then
Heat gain by the cooling water = Heat removed from the refrigerant
· ·
mw Cp (t2 – t1) = Q1 = m (H3 – H4)
The rate of water circulation in the condenser is given by
·
· m (H3 – H4)
mw = C (t – t ) , kg/s ... (6.33)
p 2 1

where t2 = Final water temperature = Temperature of water leaving the condenser in K.


t1 = Initial water temperature = Temperature of water entering the condenser in K
Cp = Heat capacity of water, kJ/(kg·K)
Chemical Engineering Thermodynamics - I 6.14 Refrigeration

The rate of work input, i.e., the power input to the compressor is given by
· ·
Power input = m W = m (H3 – H2) , kJ/s ... (6.34)
The rate of heat removal or absorption from the cold space is
· ·
m (H2 – H1) kJ/s = m (H2 – H1) × 3600 kJ/h
Therefore, the capacity of the refrigeration plant or unit is
· 1 ton
Capacity = m kg/s × (H2 – H1) kJ/kg × 3600 s/h, kJ/h × 12660 kJ/h , tons ... (6.35)

i.e.‚ Capacity = m
·
kg/s × Q2 kJ/kg × 3600 s/h‚ kJ/h × 12660 kJ/h ‚ tons
1 ton
 
Advantages of vapour compression system over the air refrigeration system :
1. The COP of the vapour compression system is quite high compared to that of the air
refrigeration system (since the vapour compression system approaches nearly to the
Carnot cycle). (Therefore, lower power requirement for a given capacity.)
2. The weight of refrigerant circulated per ton of refrigeration is less compared to that of
air circulated in the air refrigeration system.
In the vapour compression system, latent heat of vaporization of the refrigerant, which is
high, is responsible for the extraction of refrigerating effect, whereas in the air refrigeration
system, sensible heat of air, which is very low, is responsible for the extraction of refrigerating
effect. (Therefore, the weight of refrigerant is less and that of air is comparatively high.)
Disadvantages of vapour compression system compared to the air refrigeration system :
1. Very high initial investment.
2. The prevention of leakage of the refrigerant is a major problem (whereas air leakage
will not create any problem).
VAPOUR ABSORPTION REFRIGERATION SYSTEM
In a vapour compression system, electricity is the energy source for refrigeration, whereas in
a vapour absorption system, heat is the energy source for refrigeration. Therefore, the former is
an electrically operated system and the latter is a heat operated system.
The basic difference between a vapour compression system and a vapour absorption system
is in the method employed for compressing the refrigerant. The compressor of the vapour
compression system is replaced by an absorber - generator assembly in the vapour absorption
system.
The components of a vapour absorption refrigeration system are : evaporator, absorber,
pump, heat exchanger, generator, condenser and throttle valve / expansion valve.
Chemical Engineering Thermodynamics - I 6.15 Refrigeration

Heat rejected
at T1 to the Heat added
surroundings Q3 at T3
NH3
vapour Generator
NH3 Condenser
liquid Weak-hot
NH3 solution

Throttle
valve
Heat exchanger

NH3
(L+V) Pressure
reducing Strong NH3
valve
solution
Refrigerator / Evaporator
NH3
Absorber Pump
vapour
Heat Heat rejected
Q2 at T2
absorbed at T1

Fig. 6.4 : Flow diagram of vapour absorption refrigeration system


Some liquid have a very high ability of absorbing large quantities of certain vapours. For
example, water has a very high ability of absorbing ammonia vapours. The solubility of ammonia
in water at low temperatures is very high. The solution formed is termed as aqua-ammonia.
The flow diagram of a vapour absorption refrigeration system, which uses water as the
absorbent and ammonia as the refrigerant, is shown in Fig. 6.4. This system is known as the
aqua-ammonia absorption system.
Ammonia vapour from the evaporator enters the absorber, where it is absorbed in the weak
solution coming from the generator. Since absorption is exothermic, heat liberated in the
absorber is removed by circulating water. The strong solution, rich in ammonia, thus formed is
then pumped to the generator after passing through the heat exchanger. In the heat exchanger, the
strong solution is preheated by absorbing heat from the high temperature weak solution entering
the exchanger from the generator.
The ammonia vapour is generated by heating strong solution by an external source (steam or
electricity) in the generator. Due to the removal of ammonia, the solution in the generator
becomes weak and this weak high temperature solution from the generator goes to the heat
exchanger. The ammonia vapour then passes to the condenser where it is condensed to high
pressure ammonia liquid. The high pressure liquid ammonia is then allowed to pass through the
throttle valve in which by Joule-Thomson expansion the liquid ammonia is cooled by partial
evaporation. The cold ammonia vapour-liquid mixture enters the evaporator, where it is
evaporated by absorbing heat from the space to be refrigerated. The ammonia vapour coming out
Chemical Engineering Thermodynamics - I 6.16 Refrigeration

of the evaporator is almost dry and enters the absorber. This completes the cycle. (The pump
increases the pressure of the strong solution to the generator pressure. The pressure reducing
valve reduces the pressure of the weak solution to the absorber pressure.) Fig. 6.4 is also
applicable for the absorption system that operates with water as the refrigerant and a lithium
bromide solution as the absorbent.
The cyclic heat engine together with the refrigerator as shown in Fig. 6.5, is equivalent to the
vapour absorption refrigeration cycle.
T3 (Generator)
Q3

Heat engine
W
Q4
T1 (Condenser and absorber)
Q1

W Refrigerator

Q2
T2 (Evaporator)

Fig. 6.5 : A vapour absorption refrigeration cycle as


equivalent to a cyclic heat engine and a cyclic refrigerator
The minimum work required for a refrigerating effect of Q2 is
Q2 (T1 – T2)
W = T2 ... (6.36)

In order to deliver this work, a heat engine operating between T3 and T1 absorbs a quantity of
heat Q3 , where
 T3 
Q3 = W T – T  ... (6.37)
 3 1
Combining Equations (6.36) and (6.37), we get

Q3 = Q2  T  T – T 
T1 – T2 T3
 2   3 1
Q2  T2  T3 – T1
∴ = T – T   T 
Q3  1 2  3 
The ratio Q2 / Q3, which is the refrigerating effect per unit of high temperature heat is known
as the COP of an vapour absorption refrigeration cycle.

COP = T – T   T 
T2 T3 – T1
... (6.38)
 1 2  3 
Chemical Engineering Thermodynamics - I 6.17 Refrigeration

This is the maximum COP. Hence, the maximum possible COP is the product of the COP of
a refrigerator working between T1 and T2 and the thermal efficiency of a heat engine working
between T3 and T1.
COMPARISON BETWEEN VAPOUR COMPRESSION AND
VAPOUR ABSORPTION SYSTEM
Vapour compression system Vapour absorption system
1. Uses electricity - a high grade energy - 1. Uses heat - a low grade energy - as the
as the energy source for refrigeration. energy source for refrigeration.
2. Uses a compressor for vapour 2. Uses an absorber - generator assembly for
compression. vapour compression.
3. Moving parts are in the compressor. 3. Moving parts are in the pump. Therefore,
Therefore, more wear and tear, high very little wear and tear, very low cost of
maintenance cost and noisy operation. maintenance and quiet operation.
4. The COP decreases considerably with 4. The COP is not affected if the system
decrease in evaporator temperature. works at lower evaporator temperature.
5. The capacity - tons of refrigeration of the 5. The capacity of the system remains almost
system decreases rapidly with decrease in same with decrease in evaporator
evaporator temperature. temperature.
6. The COP of this system decreases with 6. The COP of this system does not vary
decrease in load. Poor performance at with variation in load. The system at
reduced loads - the COP varies with load. reduced load is almost efficient as at full
load - the COP does not vary with load.
7. Charging of refrigerant is difficult. 7. Charging of refrigerant is simple.
8. More chances of leakage of refrigerant. 8. No chance of leakage of refrigerant as
there is no compressor to cause leakage.
9. Liquid traces of refrigerant in the suction 9. Liquid traces of refrigerant in the piping at
line may damage the compressor. the exit of evaporator create no danger.
REFRIGERANTS
• The working fluid in a refrigeration unit is called a refrigerant.
• A refrigerant is any substance which absorbs heat through evaporation and loses heat
through condensation in a refrigerating system.
Desirable Properties of a Refrigerant :
An ideal refrigerant should possess the following properties / characteristics :
1. It should have low boiling point.
2. It should have low freezing point.
3. It should have high latent heat of vaporization.
Chemical Engineering Thermodynamics - I 6.18 Refrigeration

4. It should have high critical temperature and low critical pressure.


5. It should have low liquid and vapour densities.
6. It should have high vapour specific heat and low liquid specific heat.
7. It should have high thermal conductivity.
8. It should have low viscosity.
9. It should be readily available and should be cheap.
10. It should be non-flammable, non-explosive and chemically stable.
11. It should be non-corrosive.
12. It should have high miscibility with lubricating oil.
13. It should be non-poisonous, non-toxic and non-irritating (odourless).
14. It should be such that its leakage should be detectable by a simple test.
15. It should give high COP in the working temperature range.
In order to avoid the entry of air or moisture into the system, it is required to operate the
system at a pressure above atmospheric pressure (i.e., a positive pressure is required). Therefore,
the boiling point of the refrigerant used should be lower than system temperatures. The moisture
in air may freeze in the expansion valve and thus cause trouble in operation. Air entry (i.e., air
leakage) into the system also increases work of compression for a definite refrigerating effect.
Since the refrigerant must operate in the cycle above its freezing point, the freezing point of
the refrigerant must be lower than system temperatures.
The latent of vaporization of the refrigerant should be as large as possible in order to reduce
the amount of refrigerant circulated in the system for the desired refrigerating effect. This
reduces the initial cost of the refrigerant. The size of the system will also then be small and this
results in low initial cost.
For high heat transfer rates, the refrigerant used should have low viscosity and high thermal
conductivity [for low µ and high k, h (heat transfer coefficient) is high]. The high heat transfer
rates reduce the area of heat transfer in the evaporator and condenser.
The COP of the refrigerant should be high. If it is so, then less power is required to drive the
compressor per ton of refrigeration.
The size of the compressor depends on the volume of suction vapour. Reciprocating
compressors are used with refrigerants with high pressures and low suction volumes, whereas
centrifugal compressors are used with refrigerant with low pressure and high suction volumes
(high volumes of the suction vapour).
The specific heat of the refrigerant in a liquid state should be low in order to minimize the
amount of vapour formed during throttling in the expansion valve. This results in increased
refrigerating effect.
Chemical Engineering Thermodynamics - I 6.19 Refrigeration

The refrigerant should be non-flammable and non-explosive, when mixed with air from a
safety point of view - so that excessive danger will not result in cases of fires or overheated
conditions.
The critical temperature of the refrigerant should be higher than the operating temperature of
the system. If reverse is the case, i.e., if the operating temperature is higher than critical
temperature, then it will be impossible to condense the gas by compressing it to high pressures.
The critical pressure should be low in order to have a low condensing pressure.
The refrigerant should not be toxic or poisonous for the safety of operators in case of
leakage.
The condenser pressure should be low since it necessitates / requires relatively light weight
equipment and consequently lower initial cost.
Refrigerants : Inorganic chemicals such as ammonia, carbon dioxide and sulphur dioxide,
halogenated hydrocarbons (methane and ethane based) such as methyl chloride, dichloro -
difluoro methane, monochloro - difluoro methane, ethyl chloride, trichloro - trifluoro ethane, etc.
and hydrocarbons such as ethane, propane, butane, propylene and isobutane are used as
refrigerants.
Halogenated hydrocarbons such as freon-12 (R-12), freon-22 (R-22), freon-13 (R-13) are
classified as group 1. These are non-toxic and relatively non-flammable and non-explosive.
Group-2 refrigerants are toxic or flammable or both and include ammonia, methyl chloride and
sulphur dioxide. Group-3 refrigerants are highly flammable and explosive and include
hydrocarbons such as methane, ethane, ethylene, propane, propylene, butane and isobutane.
Methane-based refrigerants are denoted by a number of two digits, wherein the first digit
minus one is the number of hydrogen atoms and the second digit indicates the number of fluorine
atoms, while the other atoms in the compound are chlorine which are equal to four minus the
sum of hydrogen and chlorine atoms.
(i) The refrigerant methane (CH4) is designated as R-50.

i.e., for R-50 : The first digit is 5 and the second digit is zero (0).
We have, Number of hydrogen atoms = First digit – 1 = 5 – 1 = 4
Number of fluorine atoms = Second digit = 0
Number of chlorine atoms = 4 – Sum of hydrogen and fluorine atoms = 4 – (4 + 0) = 0
Therefore, R-50 is methane (CH4).
(ii) The refrigerant dichloro - difluoro methane (CCl2F2) is designated as R-12, i.e., R-12 is
dichloro - difluoro methane.
Chemical Engineering Thermodynamics - I 6.20 Refrigeration

For R-12 : The first digit is 1 and the second digit is 2.


∴ Number of hydrogen atoms = First digit – 1 = 1 – 1 = 0
Number of fluorine atoms = Second digit = 2
Number of chlorine atoms = 4 – Sum of hydrogen and fluorine atoms = 4 – (0 + 2) = 2
Since C = 1, H = 0, Cl = 2 and F = 2, R-12 is CCl2 F2, i.e., dichloro - difluoro methane.
(iii) The refrigerant monochloro - difluoro methane CHClF2 is designated as R-22.
For CHClF2 : Number of hydrogen atoms = 1
Number of chlorine atoms = 1
Number of fluorine atoms = 2
Number of hydrogen atoms = First digit – 1
∴ 1 = First digit – 1
∴ First digit = 1 + 1 = 2
Number of fluorine atoms = Second digit
∴ Second digit = 2
Therefore, CHClF2 is designated as R-22.
For example, Refrigerant-22 (R-22) is monochloro - difluoro methane.
For R-22 : The first digit is 2 and the second digit is also 2.
Number of hydrogen atoms = First digit – 1 = 2 – 1 = 1
Number of fluorine atoms = Second digit = 2
Number of chlorine atoms = 4 – Sum of hydrogen + fluorine atoms
= 4 – (1 + 2) = 1
Therefore, R-22 is CHClF2 - monochloro - difluoromethane.
Ethane - based compounds are denoted by a number of three digits, where the first digit is
always one, the second digit minus one is the number of hydrogen atoms, the third digit indicates
the number of fluorine atoms and all other atoms in the hydrocarbon are chlorine, which are
equal to six minus the sum of hydrogen and fluorine atoms.
(i) For example, R-113, trichloro - trifluoro ethane (C2Cl3F3).
For R-113 : Number of hydrogen atoms = Second digit – 1 = 1 – 1 = 0
Number of fluorine atoms = Third digit = 3
Number of chlorine atoms = 6 – Sum of hydrogen + fluorine atoms = 6 – (0 + 3) = 3
Since H = 0, Cl = 3, and F = 3, R-113 is CCl3F3 - trichloro - trifluoro ethane.
(ii) For example, R-142 is monochloro - difluoro ethane.
For R-142 : Number of hydrogen atoms = Second digit – 1 = 4 – 1 = 3
Number of fluorine atoms = Third digit = 2
Number of chlorine atoms = 6 – (3 + 2) = 1
Since H = 3, Cl = 1 and F = 2, R-142 is monochloro - difluoro ethane (CH3ClF2).
Chemical Engineering Thermodynamics - I 6.21 Refrigeration

LIQUEFACTION OF GASES
Gases are liquefied in order to make their storage, handling and transportation simple.
For example, liquid propane in cylinders is used as a domestic fuel. In addition, gas mixtures are
liquefied in order to separate them into their components by fractionation. For example,
liquefaction of air to obtain liquid oxygen and nitrogen.
When a gas is cooled to a temperature in the two-phase region, then the gas will be liquefied.
Liquefaction of gases can be achieved by
(i) Heat exchange at constant pressure.
(ii) Joule - Thomson expansion through a throttle valve.
(iii) Isentropic expansion in an expander in which work is produced.
HEAT EXCHANGE AT CONSTANT PRESSURE
This method requires a heat sink at a temperature lower than the temperature to which the
gas is to be cooled. It is used for precooling a gas prior to liquefaction of the gas by the other two
methods.
JOULE - THOMSON EXPANSION
For liquefaction of a gas by this method, the temperature of the gas before throttling should
be below the maximum inversion temperature. Since the maximum inversion temperature for
nearly all gases is above the normal ambient temperature, cooling can be obtained by the
Joule - Thomson expansion.
The Linde process is used for liquefaction of gases which have a positive Joule - Thomson
coefficient.
The Linde process in which Joule - Thomson effect is utilized for cooling and ultimately
liquefying air is shown in Fig. 6.6.

Fig. 6.6 : Linde liquefaction process


Chemical Engineering Thermodynamics - I 6.22 Refrigeration

After compressing the air, it is cooled in a cooler and then in a heat exchanger by counter-
current heat exchange with the liquefied air from a separator. It is then expanded in a throttle
valve where its temperature is dropped due to the Joule - Thomson effect. The lower the
temperature of the air entering the throttle valve, the greater the fraction of air that is liquefied.
The two-phase mixture of air enters the separator from where liquid air is withdrawn and
unliquefied air is recycled.
· ·
Let mf be the rate at which liquid air is produced and m be the rate at which air is
compressed.
The yield of the liquefied air is given by
·
mf
y = · ... (6.39)
m
The steady-state energy balance over the control surface-I is
· · · ·
m H2 = mf H6 + (m – mf) H8
· · · ·
m H2 = mf H6 + m H8 – mf H8
· ·
m (H2 – H8) = mf (H6 – H8)
·
mf H2 – H8
∴ y = · = H –H ... (6.40)
6 8
m
where H2, H6 and H8 are the specific enthalpies at the positions indicated in Fig. 6.6.
The energy balance around the separator, throttle valve and heat exchanger is
· · · ·
m H3 = mf H6 + (m – mf) H8 ... (6.41)
The energy balance for the compressor is
· ·
m H1 + WC = m H2 + QL ... (6.42)
where WC is the work input required and QL is the rate of heat loss from the compressor to the
surroundings.
·
WC = m (H2 – H1) + QL
WC QL
∴ · = (H 2 – H1 ) + · ... (6.43)
m m
This is the minimum work required.
Specific work consumption (W) : It is the energy required per unit mass of liquid air
produced.
·
WC m WC
W = · × · = · ×y
m mf m
Chemical Engineering Thermodynamics - I 6.23 Refrigeration

Substituting for y from Equation (6.40), we get


WC H2 – H6
W = · H – H  ... (6.44)
m  6 8

· ·
In the above equations, m and mf are in kg/s, H in kJ/kg, QL in kJ/s and WC is in kJ/s.
CLAUDE PROCESS
In the Claude process, the throttle valve in the Linde process is replaced by an expander /
expansion engine. In the expander, the gas stream does some work at the expense of its energy.
Make-up m
gas Compressor Cooler
1 2
m - mf
3
8

Heat exchanger
4
Unliquefied
air Expander WE
7
5
6
Separator Liquid
mf
Fig. 6.7 : Claude liquefaction process
An expander is a work producing device which operates with a high pressure gas as the
working fluid.
The steady-state energy balance around the separator, heat exchanger and cooler is
· · · ·
m H3 = mf H6 + (m – mf) H8 + WE ... (6.45)
where WE is the work of the expander. If it operates adiabatically, then the work produced is
given by
·
WE = – m (H5 – H4) ... (6.46)
1
[For negligible changes in KE and PE‚ 2 ∆u = 0‚ g ∆z = 0
2

For adiabatic operation, Q = 0


1
∆H + 2 ∆u2 + g ∆z = Q – WS ... First law for a steady flow process

∆H = – WS
WS = – ∆H in kJ/kg
·
∴ WE = WS = – m (H5 – H4) in kJ/s ]
Chemical Engineering Thermodynamics - I 6.24 Refrigeration

SOLVED EXAMPLES

Example 6.1 :
A Carnot refrigerator requires 1.3 kW power per ton of refrigeration in order to maintain a
temperature of – 38°C in the cold space. Find (i) COP, (ii) higher temperature of the cycle and
(iii) the heat delivered and COP when it operates as a heat pump.
Solution :
The COP of a Carnot refrigerator is given by
Heat absorbed Refrigerating effect
COP = Work required = Work required
Given : Capacity = 1 ton. This is equivalent to a refrigerating effect of 12660 kJ/h.
12660
Refrigerating effect = 12660 kJ/h = 3600 = 3.5167 kJ/s = 3.5167 kW

Work required (expressed as power) = 1.3 kW


3.5167
(COP)R = 1.3 = 2.705 ≈ 2.7 ... Ans. (i)

The COP can be expressed as


T2
(COP)R = T – T
1 2

where T2 = – 38°C = 273 – 38 = 235 K


235
2.7 = T – 235
1

235
∴ T1 – 235 = 2.7

∴ T1 = 322 K (= 49°C)
Higher temperature of the cycle = 49°C ... Ans. (ii)
Heat pump :
Heat delivered as heat pump at T2 = Heat absorbed at T1 + Work required
= 3.5167 + 1.3 = 4.8167 kW
= 4.8167 kJ/s ... Ans. (iii)
Heat delivered
(COP)HP = Work required
4.8167
= 1.3 = 3.7 ... Ans (iii)
or (COP)HP = 1 + (COP)R
= 1 + 2.7 = 3.7 ... Ans. (iii)
Chemical Engineering Thermodynamics - I 6.25 Refrigeration

Example 6.2 :
A cold storage is to maintained at – 4°C when the surroundings are at 30°C. The heat
leakage from the surroundings into the cold storage is found to be 25 kW. The actual COP of the
refrigeration plant employed is found to be one-third that of an ideal refrigeration plant working
between the same temperatures. Estimate the power required to run the plant.
Solution :
T2 = – 4°C = 269 K, T1 = 30°C = 303 K
Heat leakage into the cold storage = Q2 = 25 kW = 25 kJ/s
[This heat is to be absorbed or extracted to maintain the cold storage at – 4°C. Therefore
it is Q2.]
The COP of an ideal refrigerator is given by
T2 269
(COP)iR = T – T = 303 – 269 = 7.912
1 1

1
Given : Actual COP = 3 Ideal COP

∴ Actual COP = COP of the actual operating refrigeration plant


1
= 3 (7.912)
= 2.637 ≈ 2.64
W = Work required by the actual refrigeration plant
Q2 = Heat absorbed = Refrigerating effect = 25 kJ/s
Actual COP = 2.64
Q2
(COP)actual = W
Q2
∴ W = (COP)
actual

25
= 2.64
= 9.469 ≈ 9.47 kJ/s = 9.47 kW
The power required to run the plant = Work required per unit time = 9.47 kW ... Ans.
Example 6.3 :
A refrigerating machine working on Bell - Coleman cycle operates under the following
conditions :
(i) Temperature of the refrigerator = 150 K
(ii) Temperature of the cooler = 300 K
(iii) Pressure in the refrigerator = 1 bar
(iv) Temperature of air at the entrance of refrigerator is 40 K less than that of the
refrigerator.
Chemical Engineering Thermodynamics - I 6.26 Refrigeration

Estimate : (i) Refrigerating effect, (ii) Net work, (iii) COP of the machine, and (iv) Cooler
pressure.
Assume both compression and expansion processes to be isentropic with γ = 1.4, Cp of air
= 1.05 kJ/(kg·K).
Solution :

Q1
P2 4 3 2-3 = Compression
3-4 = Heat rejection
4-1 = Expansion
1-2 = Heat absorption
P
Q2
P1
1 2

V
Fig. E 6.3
Given : T2 = 150 K - Refrigerator temperature
T4 = 300 K - Cooler temperature
T1 = T2 – 40 = 150 – 40 = 110 K
P1 = 1 bar = Pressure in the refrigerator (at the entrance to compressor)
P2 = ? = Cooler pressure, at the exit of compressor
Cp = 1.05 kJ/(kg·K)
(i) Refrigerating effect (Q2) :
Q2 = Cp (T2 – T1), in kJ/kg
= 1.05 (150 – 110)
= 42 kJ/kg
For the cycle, we have
γ–1
T4 T3 P2 γ
= = P 
T1 T2  1
T4 T3
∴ T1 = T2

T3 = T2 T  = 150 × 110 = 409.1 K


T4 300

 1  
T3 = 409.1 K (Temperature of the air at the exit of compressor)
Chemical Engineering Thermodynamics - I 6.27 Refrigeration

γ–1
T3 P2 γ
Cooler pressure : = P 
T2  1
γ
P2 = T3γ–1
∴ P  T 
 1  2
We have : P1 = 1 bar, T3 = 409.1 K, T2 = 150 K and γ = 1.4
1.4
1.4 – 1
P2 409
1 = 150

P2 = 1 ×  150 
409.1 1.4 / 0.4
= 33.5 bar
 
Cooler pressure = 33.5 bar ... Ans (iv)
The net work required is given by
Wnet = Cp (T3 – T4) – Cp (T2 – T1) , kJ/kg
= 1.05 (409.1 – 300) – 1.05 (150 – 110)
= 72.555 ≈ 72.56 kJ/kg ... Ans. (ii)
The COP is given by
Q2 42
COP = W = 72.56 = 0.5788 ≈ 0.58 ... Ans. (iii)
net

OR : The COP of air refrigeration system operating on Bell - Coleman cycle is given by
1
COP = γ–1
P2 γ – 1
P 
 1
1
= 1.4 – 1 = 0.5789 ≈ 0.58 ... Ans. (iii)
1.4
33.5 –1
 1 
Example 6.4 :
A refrigerator of 6 tons capacity operating on Bell - Coleman cycle has an upper limit of
pressure of 5.2 bar. The pressure and temperature at the entrance of compressor are 1.0 bar and
16°C respectively. The compressed air cooled at constant pressure to a temperature of 41°C
enters the expander.
Calculate : (i) COP, (ii) rate of air circulation in kg/min, (iii) Power required to drive the
unit.
Assume compression and expansion to be isentropic with γ = 1.4.
Take Cp = 1.003 kJ/(kg·K).
Chemical Engineering Thermodynamics - I 6.28 Refrigeration

Solution :

Given : Capacity = 6 tons


P2 = 5.2 bar = Cooler pressure

P1 = 1.0 bar = Pressure at the entrance of compressor (i.e., in the evaporator)

T2 = 16°C = 289 K = Temperature of the air at the entrance of compressor

T4 = 41°C = 314 K = Temperature of the air at the entrance of expander


(leaving the cooler)

γ = 1.4 and Cp = 1.003 kJ/(kg·K)

P2 4 3 2-3 = Compression
3-4 = Heat rejection
4-1 = Expansion
1-2 = Heat absorption
P

P1
1 2

Fig. E 6.4

For isentropic compression 2 - 3 :


γ–1
γ
T3 P2
T2 = P1
γ–1 1.4 – 1
γ
T3 = T2 P 
P2 1.4
= 289  1 
5.2
= 462.87 K
 1  

For isentropic expansion 4 - 1 :


γ–1
T4 P2 γ
= P 
T1  1
1.4 – 1
1.4
314 5.2
T1 =  1 

∴ T1 = 196.05 ≈ 196 K
Chemical Engineering Thermodynamics - I 6.29 Refrigeration

The COP of an air refrigerator (for isentropic processes) is given by


T1
COP = T – T
4 1

196
= 314 – 196 = 1.66 ... Ans.
1
or COP = γ–1
P2 γ – 1
P 
 1
1
= 1.4 – 1 = 1.66 ... Ans.
1.4
5.2 –1
1
T2 289
or COP = T – T = 462.87 – 289 = 1.66 ... Ans.
3 2

(ii) Rate of air circulation :


Refrigerating effect per kg of air = the space to be refrigerated per kg of air 
Heat absorbed (i.e.‚ heat removed) from

∴ Q2 = Cp (T2 – T1)
= 1.003 (289 – 196) = 93.279 ≈ 93.28 kJ/kg
Given : Capacity of the refrigerator = 6 tons
We know that 1 ton of refrigeration is equivalent to a refrigeration rate (or rate of heat
removal) of 12660 kJ/h.
Refrigerating effect produced by the refrigerator
 per hour or rate of refrigeration/rate of heat  = 6 × 12660 = 75960 kJ/h
 removal from the refrigerator 
Rate of heat removal/refrigeration  Heat removed 
 = or refrigerating ×  rate in kg/h 
Air circulation
 from the refrigerator or
 refrigerating effect in kJ/h   effect in kJ/kg 
75960
∴ Rate of air circulation = 93.28 = 814.32 kg/h = 13.57 kg/min ... Ans. (ii)
Power required to drive the unit :
Refrigerating effect produced by the refrigerator
COP = Work required to produce the effect
75960‚ kJ/h
1.66 = Net work required‚ kJ/h
75960
∴ Net work required = 1.66 = 45759 kJ/h
45759
= 3600 = 12.71 kJ/s = 12.71 kW
∴ Power required to drive the unit = 12.71 kW ... Ans. (iii)
Chemical Engineering Thermodynamics - I 6.30 Refrigeration

OR, When we use Q2 and Wnet based on per kg of air, then


Q2
COP = W
net
Q2 93.28
∴ Wnet = COP = 1.66 = 56.19 kJ/kg, since Q2 is in kJ/kg

Power = Work done per unit time


· ·
Power = m · Wnet , where m = rate of air circulation
= 814.32 kg/h × 56.19 kJ/kg
= 45749.898 ≈ 45750 kJ/h
45750
= 3600 = 12.708 ≈ 12.71 kJ/s = 12.71 kW ... Ans. (iii)

Example 6.5 :
A Bell - Coleman refrigerator operates between 1 bar and 6 bar. Air is drawn from the cold
chamber at – 5°C, compressed according to the polytropic law PV1.25 = C and is cooled to 37°C
before entering the expander. The expansion is isentropic.
Calculate : (i) COP, (ii) refrigeration capacity in tons and (iii) rate of air circulation if
500 kg of ice is produced per day at 0°C when water is supplied at 20°C.
Data : γ = 1.4, Cp of air = 1.0 kJ/(kg·K), Cp of water = 4.1868 kJ/(kg·K) and latent heat
of ice = 335 kJ/kg.
Solution :

Fig. E 6.5
γ
Given : Expansion is isentropic, i.e., it follows the law PV = C, whereas compression
follows the law PVn = C, where n = 1.25.
T2 = – 5oC = – 5 + 273 = 268 K - Temperature of the air before compression
T3 = Temperature of the air after compression
Chemical Engineering Thermodynamics - I 6.31 Refrigeration

For a polytropic compression :


n–1
T3 P2 n ... for the process 2 - 3
= P 
T2  1
where n = 1.25, P1 = 1 bar = Pressure in the refrigerator
P2 = 6 bar = Pressure in the cooler
1.25 – 1
T3 6
1.25

268 = 1
∴ T3 = 383.5 K
T4 = Temperature of the air leaving the cooler = 37°C = 310 K
Heat removed or absorbed per day for making 500 kg of ice from water at 293 K
= mCp (∆T) + mλ
where m = 500 kg, λ = 335 kJ/kg, Cp = 4.1868 kJ/(kg·K) and ∆T = 293 – 273 K = 20 K
Heat removal per day = 500 × 4.1868 × 20 + 500 × 335
= 209368 kJ
Rate of heat removal = 209368 kJ/day
= 8723.67 kJ/h

Heat removal in kJ/h =  in kJ/kg  × 


Heat removal Rate of air circulation
in kg/h 
8723.67 kJ/h
Rate of air circulation = 82.2 kJ/kg

= 106.127 kg/h = 1.768 ≈ 1.77 kg/min ... Ans. (iii)


Refrigerator capacity in tons of refrigeration : 1 ton of refrigeration is equivalent to a rate
of heat removal or a refrigeration rate of 12660 kJ/h.
Rate of heat removal in the given refrigerator = 8723.67 kJ/h
1
Capacity of the refrigerator = 8723.67 × 12660
= 0.689 ≈ 0.70 ton ... Ans. (ii)
The process 4-1 (expansion) is isentropic. For this process,
γ–1
T4 P2 γ
= P 
T1  1
1.4 – 1
1.4
310 6
T1 = 1
∴ T1 = 185.799 ≈ 185.8 K
Chemical Engineering Thermodynamics - I 6.32 Refrigeration

The net work required per kmol of air is given by


Wnet = WC – WE
n γ
= n – 1 R (T3 – T2) – R (T4 – T1)
γ–1
where R is 8.31451 kJ/(kmol·K).
Molecular weight of air = 29 kg/kmol
The net work required per kg of air is
nR γR
Wnet = 29 (n – 1) (T3 – T2) – (T – T1)
29 (γ – 1) 4
where n = 1.25, γ = 1.4, T3 = 385.5 K, T2 = 268 K, T4 = 310 K and T1 = 185.8 K
Substituting the values,
1.25 × 8.31451 1.4 × 8.31451
Wnet = 29 (1.25 – 1) (385.5 – 268) – 29 (1.4 – 1) (310 – 185.8)

= 168.44 – 124.63 = 43.81 kJ/kg


Heat removed from the cold chamber (refrigerating effect) per kg of air is
Q2 = Cp (T2 – T1)
Q2 = 1 × (268 – 185.8) , Cp = 1 kJ/(kg·K)
= 82.2 kJ/kg
The COP of the refrigerator is
Q2
COP = W
net
82.2
COP = 43.81 = 1.876 ≈ 1.88 ... Ans. (i)
Given : We have to produce 500 kg of ice per day at 0°C (273 K) from water at
20°C (293 K).
In order to produce ice, we have to remove (i) sensible heat from water to bring down its
temperature from 293 K to 273 K and (ii) latent heat to convert water at 273 K to ice at 273 K.
Example 6.6 :
A refrigerating unit using freon-12 (R-12) as the working fluid operates between 18°C and
37°C. The rate of circulation of refrigerant is 2 kg/min and the efficiency of the compressor is
0.85. Using the following data of enthalpy, calculate :
(i) the capacity of the plant in tons of refrigeration.
(ii) the power required to run the unit and
(iii) the COP of the unit.
Data : The enthalpies of R-12 liquid at 37°C is 455 kJ/kg. The enthalpies of R-12 entering
and leaving the compressor are 563.15 kJ/kg and 595.4 kJ/kg respectively.
Chemical Engineering Thermodynamics - I 6.33 Refrigeration

Solution :

4 3
3'
1-2 : Heat absorption at constant P
2-3 : Isentropic compression
P
2 3-4 : Heat rejection at constant P
1
4-1 : Adiabatic expansion at constant H

H
Fig. E 6.6
The process of expansion 4-1 occurs at constant enthalpy. So H4 = H1.
H2 = Enthalpy of R-12 entering the compressor = 563.15 kJ/kg
H3 = Enthalpy of R-12 leaving the compressor = 595.4 kJ/kg
H4 = Enthalpy of R-12 leaving the condenser = 455 kJ/kg
ηC = Compressor efficiency = 0.85
·
m = Rate of circulation of R-12 = 2 kg/min = 0.0333 kg/s
H4 = H1 = 455 kJ/kg
1 ton of refrigeration is equivalent to a heat removal or refrigeration rate of 12660 kJ/h.
The capacity of the plant in tons of refrigeration is given by
· 1 ton
Capacity = m kg/s × (H2 – H1) kJ/kg × 3600 s/h, kJ/h × 12660 kJ/h , tons

1
= 0.0333 × (563.15 – 455) × 3600 × 12660

= 1.027 ≈ 1.03 tons ... Ans. (i)


Power required to run the unit :
·
W = WC = m kg/s × (H3 – H2) kJ/kg
= 0.0333 × (595.4 – 563.15)
= 1.074 kJ/s ... for 100% efficient compressor
But the compressor efficiency is 0.85 or 85%.
∴ The power required to run the unit is
100
= 1.074 × 85 = 1.263 kJ/s = 1.263 kW ... Ans. (ii)
Chemical Engineering Thermodynamics - I 6.34 Refrigeration

The COP of a vapour compression cycle is given by


H2 – H1 H2 – H4
COP = H – H = H – H , since H1 = H4
3 2 3 2

563.15 – 455
= 595.4 – 563.15

= 3.35 ... for 100% efficient compressor


∴ COP of the unit = ηC × COP of the cycle
= 0.85 × (3.35) = 2.8475 ≈ 2.85 ... Ans. (iii)
Refrigerating effect in kJ/s
OR COP of the unit = Power required to run the plant in kJ/s

·
m kg/s × (H2 – H1) kJ/kg
= 1.2630 kJ/s
0.0333 × (563.15 – 455)
= 1.263 = 2.85 ... Ans. (iii)

Example 6.7 :
A vapour compression machine using R-12 as the refrigerant is employed to maintain a
temperature of – 23°C in the refrigerated space. The ambient temperature is 37°C. A minimum
10°C temperature difference is needed in the evaporator and condenser for heat transfer. There
is no subcooling of liquid R-12 in the condenser and R-12 enters the compressor as a dry
saturated vapour. Estimate (i) tonnes of refrigeration, (ii) power requirement and (iii) the ratio
of COP of this cycle to the COP of a Carnot cycle if the refrigerant flow rate is 1 kg/min.
Data : Enthalpy of saturated vapour entering the compressor = 336.8 kJ/kg
Enthalpy of superheated vapour leaving the compressor = 384 kJ/kg
Enthalpy of saturated liquid refrigerant leaving the condenser = 245.7 kJ/kg
Solution :
Refrigerator temperature / Refrigerated space temperature,
T1 = – 23°C = – 23 + 273 = 250 K
Ambient temperature, T2 = 37°C = 310 K
The COP of a Carnot cycle is given by
T1 250
(COP)Carnot = T – T = 310 – 250 = 4.1666 = 4.167
2 1

Given : 10°C temperature difference is required in the condenser and evaporator for heat
transfer.
Therefore, the temperature of the refrigerant in the evaporator must be less than T1 by 10°C.
∴ Temperature of R-12 in the evaporator = 250 – 10 = 240 K
Chemical Engineering Thermodynamics - I 6.35 Refrigeration

Similarly, the temperature of R-12 in the condenser must be higher than ambient temperature
by 10°C from heat transfer from R-12 to ambient air.
∴ Temperature of R-12 in the condenser = T2 + 10 = 310 + 10 = 320 K

Q1 3
4
320 320 K
T

310 4 3

250 P
240 1 2
1 2 240 K
Q2

S H
Fig. E 6.7
1 - 2 : Heat absorption at constant P, 2 - 3 : Isentropic compression,
3 - 4 : Heat rejection at constant P, 4 - 1 : Adiabatic expansion at constant H
We have, H4 = H1 (Refer to the figure given above)
H2 = Enthalpy of dry saturated vapour entering the compressor = 336.8 kJ/kg
H3 = Enthalpy of superheated vapour leaving the compressor = 384 kJ/kg
H4 = H1 = Enthalpy of saturated liquid refrigerant leaving the condenser = 245.7 kJ/kg
The COP of a vapour compression cycle is given by
H2 – H1
(COP)cycle = H – H
3 2

336.8 – 245.7
= 384 – 336.8 = 1.93

(COP)cycle
Relative COP = (COP)
Carnot

1.93
= 4.167 = 0.463 or 46.3% ... Ans. (iii)
· 1
Rate of refrigerant circulation = m = 1 kg/min = 60 = 0.0167 kg/s
The capacity of the unit in tons of refrigeration is given by
· 1 ton
Capacity = m kg/s × (H2 – H1) kJ/kg × 3600 s/h, kJ/h × 12660 kJ/h , tons
1
= 0.0167 × (336.8 – 245.7) × 3600 × 12660

= 0.4326 ≈ 0.433 tons ... Ans. (i)


Chemical Engineering Thermodynamics - I 6.36 Refrigeration

Power required by the unit is given by


· · ·
Power = m W = m WC = m kg/s × (H3 – H2) kJ/kg , kJ/s
= 0.0167 × (384 – 336.8)
= 0.788 ≈ 0.79 kJ/s = 0.79 kW ... Ans. (ii)
Example 6.8 :
A refrigerating machine with ammonia as the refrigerant is used for producing 500 kg/h of
ice from water. Ammonia boils at 266 K and condenses at 293 K. The water in the condenser is
thereby heated from 283 K and 288 K. Calculate the theoretical minimum power required by the
compressor and the circulation rate of water.
Data : The latent heat of fusion of water is 339.1 kJ/kg
Solution :
·
Production rate of ice = m = 500 kg/h
Latent heat of fusion of water = λ = 339.1 kJ/kg
Heat absorbed during vaporization in the refrigerator = QL
·
QL = Q2 = m λ
= 500 × 339.1 = 169550 kJ/h
T2 = TL = Temperature at which ammonia boils = 266 K
T1 = TH = Temperature at which ammonia condenses = 293 K
The COP of an ideal refrigerator is given by
TL T2
COP = T – T = T – T
H L 1 2

266
= 293 – 266 = 9.852 ... maximum COP

The work required will be minimum for an ideal refrigerator having maximum COP.
QL
COP = W
QL 169550
∴ W = COP = 9.852

= 17209.7 kJ/h = 4.78 kJ/s = 4.78 kW


The theoretical minimum power required by the compressor is 4.78 kW
Heat rejected in the condenser = Q1
We have, Q1 – Q2 = W
∴ Q1 = Q2 + W , where W = 4.78 kW = 17209.7 kJ/h
∴ Q1 = 169550 + 17209.7 = 186759.7 kJ/h
Chemical Engineering Thermodynamics - I 6.37 Refrigeration

.
Let mw be the circulation rate of water.
T1 = 283 K and T2 = 288 K ... for water
Cp of water = 4.187 kJ/(kg·K)
In the condenser heat is rejected to the water thereby it gets heated from 283 K to 288 K.
.
∴ QH = Q1 = mwCp (T2 – T1)
.
186759.7 = mw × 4.187 (288 – 283)
.
∴ mw = 8920.9 kg/h = 8.9209 × 103 kg/h ...
... Ans.
Example 6.9 :
An ideal vapour compression unit employing Freon-12 operates between an evaporator
temperature of 243 K and a condenser temperature of 308 K. Determine the refrigeration
capacity (in tons of refrigeration) and the rate of circulation of the refrigerant if the power input
to the compressor is 50 kW.
Data : Enthalpy of saturated liquid Freon-12 at 308 K = 69.55 kJ/kg
Enthalpy of saturated vapour at 243 K = 174.2 kJ/kg
Enthalpy of superheated vapour leaving the compressor = 200 kJ/kg
Solution :
H2 = Enthalpy of saturated vapour at 243 K = 174.2 kJ/kg
H1 = Enthalpy of saturated liquid at 308 K = 69.55 kJ/kg
H3 = Enthalpy of superheated vapour leaving the compressor = 200 kJ/kg

4 3 1-2 = Evaporation
2-3 = Compression
3-4 = Condensation
P
2 4-1 = Expansion / Throttling
1

H
Fig. E 6.9
The COP is given by
Q2 H2 – H1 174.2 – 69.55
COP = W = H – H = 200 – 174.2 = 4.056
3 2
Chemical Engineering Thermodynamics - I 6.38 Refrigeration

Power input to the compressor = W = 50 kW = 50 kJ/s


QL Q2
COP = W = W

∴ Q2 = QL = Heat absorbed = COP × W


= 4.056 × 50 = 202.8 kJ/s, since W is in kJ/s
Q2 = 202.8 × 3600 = 730080 kJ/h
For a refrigerator rated at 1 ton : QL = Q2 = 12660 kJ/h
[One ton of refrigeration is equivalent to a refrigeration rate or heat absorption rate of
12660 kJ/h.]
Refrigerator capacity in tons of refrigeration is
730080
Refrigeration capacity = 12660 = 57.668 ≈ 57.67 ton ... Ans.

OR : The capacity of the refrigerator is


· 1 ton
Capacity = m kg/s × (H2 – H1) kJ/kg × 3600 s/h, kJ/h × 12660 kJ/h , tons

The rate of heat removal from the cold space is given by


·
Q2 in kJ/s = m in kg/s × (H2 – H1) in kJ/kg … (1)
·
where m is the rate of circulation of refrigerant.
Thus, the capacity equation becomes
1 ton
Capacity = Q2 in kJ/s × 3600 in s/h, kJ/h × 12660 kJ/h , tons
We have, Q2 = 202.8 kJ/s
1
∴ Capacity = 202.8 × 3600 × 12660 = 57.668 ≈ 57.68 ton … Ans.

The rate of circulation of the refrigerant is


· Q2 in kJ/s
m = (H – H ) in kJ/kg … from Equation (1)
2 1

202.8
= (174.2 – 69.55)

= 1.9378 ≈ 1.94 kg/s … Ans.


Example 6.10 :
A cold storage plant is needed to store 20 tonnes of fish. The temperature of the fish when
supplied is 25°C and storage temperature of fish required is – 8°C.
If the cooling is achieved in 8 h, find out :
(i) the capacity of the refrigerating plant,
Chemical Engineering Thermodynamics - I 6.39 Refrigeration

(ii) the Carnot cycle COP in the temperature range provided and
(iii) the power required to run the plant if the actual COP is one-third of the Carnot COP.
Data : Specific heat of fish above the freezing point = 2.93 kJ/(kg·°C)
Specific heat of fish below freezing point = 1.25 kJ/(kg·°C)
Freezing point of fish = – 3°C
Latent heat of fish = 232 kJ/kg
Solution :
T1 = Initial fish temperature = 25°C
T2 = Freezing point of fish = – 3°C
T3 = Storage temperature = – 8°C
Cp1 = Cp of fish above – 3°C = 2.93 kJ/(kg·°C)
Cp2 = Cp of fish below – 3°C = 1.25 kJ/(kg·°C)
λ = Latent heat of fish = 232 kJ/kg
Cooling is achieved in 8 h.
The heat removed in 8 h from each kg of fish is
Q2 = 1 × Cp1 (T1 – T2) + 1 × λ + 1 × Cp2 (T2 – T3)
 Sensible heat  Latent heat  
Sensible heat 
= removed to cool + removed at T + removed to cool 
  
fish from T to T   2 fish from T to T 
 1 2  2 3

= 1 × 2.93 [25 – (– 3)] + 1 × 232 + 1 × 1.25 [– 3 – (– 8)]


= 320.29 kJ ... from 1 kg
Q2 = Heat removed from fish = 320.29 kJ/kg
Quantity of fish = 20 tonnes = 20 × 103 kg
Heat removed by the plant = 20 × 103 × 320.29 = 6405800 kJ
This heat is removed from the fish in 8 h by the plant.
6405800
∴ Rate of heat removal by the plant = Q2 = 8 = 800725 kJ/h

Rate of heat removal of 1 ton capacity plant is 12600 kJ/h.


1
∴ Capacity of the refrigerating plant = 800725 × 12660 = 63.25 ton ... Ans. (i)

OR : The capacity of the refrigerating plant is


· 1 ton
Capacity = m kg/s × (H2 – H1) kJ/kg × 3600 s/h, kJ/h × 12660 kJ/h , tons
Chemical Engineering Thermodynamics - I 6.40 Refrigeration

We have,
·
Q2 in kJ/s = m kg/s × (H2 – H1) kJ/kg
·
Q2 in kJ/h = m kg/s × (H2 – H1) kJ/kg × 3600 s/h
·
∴ m (H2 – H1) × 3600 in the capacity equation can be substituted by Q2 in kJ/h.
1 ton
∴ Capacity = Q2 in kJ/h × 12660 kJ/h

1
= 800725 × 12660 = 63.24 ton … Ans.

The Carnot COP between T1 and T2 is


T2
COP = T – T
1 2

where T2 = – 8°C = 273 + (– 8) = 265 K


T1 = 25°C = 273 + 25 = 298 K
265
(COP)Carnot = 298 – 265 = 8.03 ... Ans. (ii)
1
Given : Actual COP = 3 (Carnot COP)
1
∴ Actual COP = 3 × 8.03 = 2.677
Refrigerating effect
COP = Work input
Refrigerating effect
Work input = COP
800725
= 2.677 = 299112.8 kJ/h
Power required to run the plant 299112.8
(i.e.‚ work input in kJ/s) = 3600

= 83.08 ≈ 83 kJ/s = 83 kW ... Ans. (iii)


Example 6.11 :
A vapour - compression refrigerator employing ammonia is rated at 5 ton. The evaporator is
at 273 K and the condenser is at 303 K. The saturation pressures of ammonia corresponding to
these temperatures are 4.29 bar and 11.67 bar, respectively. The permissible temperature rise
for cooling water is 10 K.
Find : (i) The coefficient of performance,
(ii) The refrigerant circulation rate,
(iii) The minimum horse power required,
(iv) The circulation rate of cooling water.
Chemical Engineering Thermodynamics - I 6.41 Refrigeration

T (K) Ps (bar) HL (kJ/kg) HV (kJ/kg)


273 4.29 168 1300
303 11.67 300 1327
The enthalpy of superheated vapour leaving the compressor at 11.67 bar is 1445 kJ/kg and
Cp of water is 4.187 kJ/(kg.K).
Solution :

4 3 1-2 = Evaporation
2-3 = Compression
3-4 = Condensation
P
2 4-1 = Expansion / Throttling
1

H
Fig. E 6.11
(i) To find the COP : The COP of the refrigerator is
Q2 QL H2 – H1
COP = W = W = H – H
3 2

where H2 = Enthalpy of saturated vapour at 273 K = 1300 kJ/kg


H1 = Enthalpy of saturated liquid at 303 K = 300 kJ/kg
H3 = Enthalpy of superheated vapour leaving the compressor = 1445 kJ/kg
1300 – 300
∴ COP = 1445 – 1300

= 6.8965 ≈ 6.9 ... Ans. (i)


Q2 = QL = Heat absorbed
1 ton of refrigeration + Heat absorption at a rate of 12660 kJ/h
The refrigerator capacity is 5 ton ... Given.
∴ Q2 = QL = Heat absorbed per hour in the refrigerator of 5 ton capacity
12660
= 5× 1 = 63300 kJ/h

(ii) To find the minimum power required to drive the refrigerator :


Q2
We have, COP = W
Q2
∴ W = COP
Chemical Engineering Thermodynamics - I 6.42 Refrigeration

63300
Minimum power required = W = 6.9
= 9173.91 kJ/h = 9173.91 × 103 J/h
9173.91 × 103
= 3600 = 2548.31 J/s
= 2548.31 W [1 J/s = 1 W]
1 hp
= 2548.31 W × 747 W
= 3.411 ≈ 3.41 hp ... Ans. (iii)
·
(iii) To find the refrigerant circulation rate (m) :
Heat absorbed in the evaporator = Q2 = QL = 63300 kJ/h
We have, heat absorbed during vaporization = Change in the enthalpy of refrigerant
·
Q2 = QL = m (H2 – H1)
where H2 = Enthalpy of saturated vapour at 4.29 bar (273 K) = 1300 kJ/kg
H1 = Enthalpy of saturated liquid at 11.67 bar (303 K) = 300 kJ/kg
·
63300 = m (1300 – 300)
·
∴ m = 63.3 kg/h
The refrigerant circulation rate is 63.3 kg/h ... Ans (ii)
·
(iv) To find the circulation rate of cooling water (m w) : In the condenser, heat is rejected
to the water and thereby it gets heated.
Permissible rise of water temperature = ∆T = 10 K
Heat rejected in the condenser = Q1 = Q2 + W = Heat gained by water
We have : W = 3.4 hp = 2548.31 W = 9173.91 kJ/h
∴ Q1 = 63300 kJ/h + 9173.91 kJ/h
Q1 = QH = 72473.91 kJ/h
. .
∴ Q1 = QH = (mCp)w ∆T = mw Cpw ∆T
.
72473.91 = mw × 4.187 × 10
.
∴ mw = 1730.926 kg/h ≈ 1731 kg/h ... Ans. (iv)
OR : The rate of heat removal in the condenser is
· .
Q1 = m (H3 – H4) = mw Cpw ∆T
Heat given out by refrigerant = Heat gained by water
·
. m (H3 – H4)
∴ mw =
Cpw · ∆T
·
m (H3 – H1)
= , since H4 = H1
Cpw ∆T
Chemical Engineering Thermodynamics - I 6.43 Refrigeration

Substituting the values,


. 63.3 (1445 – 300)
mw =
4.187 × 100
= 1731 kg/h … Ans. (iv)
Example 6.12 :
A refrigerator, which uses R-134 as the working fluid, operates on an ideal vapour
compression cycle between 0.14 MPa and 0.80 MPa. Find : (i) the rate of heat removal from the
refrigerated space, (ii) the COP, (iii) the power input to the compressor and (iv) the heat
rejected in the condenser, if the mass flow of the refrigerant is 0.06 kg/s.
Data : Enthalpy of saturated vapour at 0.14 MPa = 236 kJ/kg
Enthalpy of saturated liquid at 0.14 MPa = 93.42 kJ/kg
Enthalpy of superheated vapour at 0.80 MPa = 272.05 kJ/kg
Solution :
Q1

H4
Expansion Condenser
valve 0.80 MPa
4 3
H3
In P or
P 1 2
H1 H2
Evaporator Compressor

Q2 W
H
Space to be cooled
or 1-2 = Evaporation
refrigerated 2-3 = Compression
3-4 = Condensation
4-1 = Throttling
(a) (b)
Fig. E 6.12
For a compressor,
H2 + W = H3 ... by the first law of thermodynamics
∴ W = H3 – H2 , kJ/kg
For a condenser,
H3 = Q1 + H4 ... by the first law of thermodynamics
∴ Q1 = H3 – H1 , since H4 = H1 for a throttling process (expansion valve)
For an evaporator,
H1 + Q2 = H2 , kJ/kg
∴ Q2 = H2 – H1
= Heat absorbed in the refrigerator by the refrigerant from the
surroundings = refrigerating effect
Chemical Engineering Thermodynamics - I 6.44 Refrigeration

The COP of a refrigerator is given by


Q2 H2 – H1
COP = W = H – H
3 2

The heat removed / absorbed from the refrigerated space is


Q2 = H2 – H1
= Change in the enthalpy of the refrigerant
= Heat absorbed from the surroundings, which gets cooled
where H1 = Enthalpy of saturated liquid at 0.14 MPa = 93.42 kJ/kg
H2 = Enthalpy of saturated vapour at 0.14 MPa = 236 kJ/kg
Q2 = 236 – 93.42
= 142.58 kJ/kg
Let us obtain the rate of heat removal, Q2.
·
m = Mass flow of refrigerant = 0.06 kg/s
·
∴ Q2 = m kg/s × (H2 – H1) kJ/kg
= 0.06 × 142.58 = 8.5548 ≈ 8.55 kJ/s = 8.55 kW
The rate of heat removal from the refrigerated space is 8.55 kW ... Ans.
Power input of the compressor :
Work required for the compressor = W = H3 – H2
where H3 = Enthalpy of superheated vapour = 272.05 kJ/kg
∴ W = 272.05 – 236 = 36.05 kJ/kg
The rate of work input, i.e., the power input to the compressor is
· ·
Power input = m in kg/s (H3 – H2) in kJ/kg = m kg/s × W kJ/kg
= 0.06 × 36.05 = 2.163 ≈ 2.16 kJ/s = 2.16 kW ... Ans.
The COP of the refrigerator is
Q2 8.55
COP = W = 2.16 = 3.958 = 3.96 ... Ans.
H2 – H1 236 – 93.42
or COP = H – H = 272.05 – 236 = 3.955 ≈ 3.96 ... Ans.
3 2

Q1 = Heat rejected in the condenser


= H3 – H1
= 272.05 – 93.42 = 178.63 kJ/kg
The rate of heat removal in the condenser is
·
Q1 = m (H3 – H1) = 0.06 kg/s × 178.63 kJ/kg
= 10.7178 ≈ 10.72 kJ/s = 10.72 kW ... Ans.
❐❐❐

You might also like