A. K. Saxena - C. M. Tiwari - Heat and Thermodynamics-Alpha Science International (2014)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 324

Heat and

Thermodynamics
Heat and
Thermodynamics

A.K. Saxena
C.M. Tiwari

a
Alpha Science International Ltd.
Oxford, U.K.
Heat and Thermodynamics
328 pgs. | 145 figs. | 09 tbls.

A.K. Saxena
C.M. Tiwari
Department of Physics
A.P.S. University
Rewa (M.P.)

Copyright © 2014
ALPHA SCIENCE INTERNATIONAL LTD.
7200 The Quorum, Oxford Business Park North
Garsington Road, Oxford OX4 2JZ, U.K.

www.alphasci.com
ISBN 978-1-84265-902-1
E-ISBN 978-1-78332-059-2
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system,
or transmitted in any form or by any means, electronic, mechanical, photocopying, recording
or otherwise, without prior written permission of the publisher.
Preface

This book has been written to cover the syllabus of B.Sc. Physics (and chemistry)
of all Colleges and Universities. The subject matter includes Basic Ideas
(Temperature and heat, measurement of temperature, gas laws and gas equation,
intensive and extensive properties, thermodynamic system and coordinator, PvT
surfaces, heat capacity, mechanical equivalent of heat and thermal expansion),
Kinetic theory (Brownian motion, kinetic model of gases and deductions of gas
laws, degrees of freedom, principle of equipartition of energy, Maxwell-Boltzmann
distribution law and its experimental verification, mean free path, transport
phenomena and thermal conductivity), Real gases and their liquefaction and
production and measurement of very low temperatures (van der Waals equation,
Joules expansion, Joule-Thomson cooling, porous plug experiment, adiabatic
demagnetization, liquefaction of gases and approach to absolute zero), The first
law of thermodynamics and thermodynamic properties, application of the first law
and concept of a perfect gas, reversible and irreversible processes and the second
and third laws of thermodynamics, Carnots’ cycle, thermodynamic temperature
scale, entropy, real heat engines, negative temperatures, thermodynamic potentials
and Maxwell’s relations and change of phase and, in the last thermal radiation
(Emission and absorption, Prevost theory and laws of radiation).
The authors are thankful to Mr. N.K. Mehra and Mr. Shashikant (Narosa
Publishing House) for bringing out the book in a short time. We hope that the
book would be helpful to undergraduate students and aid to their understanding
of the subject.

A.K. Saxena
C.M. Tiwari
Contents

Preface v
1. Basic Ideas 1.1
1.1 Thermodynamics 1.1
1.2 Zeroeth Law of Thermodynamics and Temperature 1.1
1.3 Temperature and Heat 1.1
1.4 Gas Thermometers 1.3
1.5 Absolute Zero 1.5
1.6 Triple-Point Cell 1.6
1.7 Temperature Measurement with a Constant-Volume Gas 1.8
Thermometer
1.8 International Practical Temperature Scale 1.9
1.9 Resistance Thermometer 1.10
1.10 Thermocouple 1.11
1.11 Ideal Gases and Gas Laws 1.12
1.12 Avogadro’s Number 1.15
1.13 Universal Gas Constant, Boltzmann Constant and 1.16
Ideal Gas Equation
1.14 The “System” 1.17
1.15 Intensive and Extensive Properties 1.18
1.16 Thermodynamic System and Thermodynamic Coordinates 1.19
1.17 P-v-T Surface for an Ideal Gas and Boyle’s Law 1.20
1.18 Behaviour of Real Gases (Equation of State and 1.22
P-v-T Surfaces)
1.19 Heat Capacity 1.26
1.20 Heat: Its Nature and Units 1.27
1.21 Mechanical Equivalent of Heat 1.28
1.22 Thermal Expansion 1.30
Questions 1.36
Objective Type Questions 1.37
viii Contents

2. Kinetic Theory 2.1


2.1 Brownian Motion 2.1
2.2 Theories of Brownian Motion 2.2
2.3 Vertical Distribution of Brownian Particles and 2.6
Determination of Avogadro’s Number
2.4 Ideal Gas (Macroscopic Description) 2.7
2.5 Kinetic Model of Gases 2.8
2.6 Some Deductions from Kinetic Theory 2.12
2.7 The Internal Energy and Degrees of Freedom 2.16
2.8 Principle of Equipartition of Energy 2.18
2.9 Maxwell-Boltzmann Law of Distribution of Velocities 2.20
in an Ideal Gas
2.10 Maxwell Boltzmann’s Energy Distribution Law 2.29
2.11 Experimental Verification of Maxwell’s Velocity 2.33
Distribution Law
2.12 Mean Free Path 2.36
2.13 Transport Phenomena 2.38
2.14 Gas Viscosity 2.39
2.15 Thermal Conductivity 2.42
2.16 Gas Self Diffusion 2.43
Questions 2.52
Objective Type Questions 2.53
3. Real Gases, Liquefaction of Gases and Production
and Measurement of Very Low Temperatures 3.1
3.1 Perfect Gas and Real Gas 3.1
3.2 Experiments on the Behaviour of Real Gases 3.2
3.3 Critical Point 3.6
3.4 Continuity of State 3.7
3.5 Form of Equation of State 3.7
3.6 van der Waals Equation and the Critical Point 3.8
3.7 Comparison between Experimental and Theoretical 3.11
P-V Curves
3.8 van der Waals Equation and the Boyle Temperature 3.12
3.9 Corresponding States 3.14
3.10 Joule’s Expansion of an Ideal Gas and van der Waal’s Gas 3.15
3.11 Joule’s Coefficient for Real Gas 3.16
Contents ix

3.12 Joule-Thomson Cooling 3.17


3.13 Temperature of Inversion 3.21
3.14 Principle of Joule Thomson’s Porous Plug Experiment 3.22
3.15 Change in Temperature in an Adiabatic Change 3.27
3.16 Distinction between Joule’s Expansion, Joule-Thomson’s 3.29
Expansion and Adiabatic Expansion
3.17 Adiabatic Demagnetisation 3.29
3.18 Liquefaction of Gases and Approach to Absolute Zero 3.35
3.19 Production of Very Low Temperatures 3.41
3.20 Measurement of Very Low Temperatures 3.44
Questions 3.47
Objective Type Questions 3.48
4. The First Law of Thermodynamics 4.1
4.1 Introduction 4.1
4.2 Work Depends on the Path (Indicator Diagram) 4.1
4.3 Heat Depends on the Path 4.4
4.4 First Law of Thermodynamics and Internal Energy 4.5
4.5 Work Done by a Thermodynamic System in Expansion 4.8
Against External Pressure
4.6 Proof for the First Law of Thermodynamics 4.13
4.7 The Thermodynamic Property ‘Enthalpy’ 4.14
4.8 Heat Capacity 4.15
4.9 Specific Heat Capacity 4.16
4.10 Relationship between Cp and Cv 4.17
4.11 Applications of the First Law of Thermodynamics 4.23
to an Ideal Gas
4.12 Impossibility of a Perpetual-Motion Machine 4.26
4.13 Surfaces 4.27
4.14 Work in Changing the Area of a Surface Film 4.28
4.15 Paramagnetic Rod 4.29
4.16 Magnetic Work 4.30
4.17 Partial Derivatives 4.32
4.18 Concept of a Perfect Gas 4.35
Questions 4.37
Objective Type Questions 4.38
x Contents

5. The Second Law and Third Law of Thermodynamics 5.1


5.1 Reversible and Irreversible Processes 5.1
5.2 Significance of Reversible Processes 5.2
5.3 Need for Formulations of Second Law of Thermodynamics 5.3
5.4 The Second Law of Thermodynamics 5.4
5.5 Heat Engine and its Efficiency 5.6
5.6 Carnot’s Engine: Carnot’s Cycle and its Efficiency 5.7
5.7 Working of a Heat Engine/Refrigerator Operating 5.13
on a Carnot Cycle
5.8 Demonstration of the Fact that the Carnot Cycle is the 5.15
Most Efficient Cycle Operating between Two
Fixed-Temperature Reservoirs
5.9 Thermodynamic Temperature Scale 5.17
5.10 Relation between Thermodynamic Scale, Celsius Scale, 5.21
Rankine Scale and the Fahrenheit Scale of Temperatures
5.11 The Clausius Theorem 5.23
5.12 The Clausius Inequality 5.24
5.13 Entropy 5.26
5.14 Physical Significance of Entropy 5.27
5.15 Thermodynamic Definition of Temperature 5.28
5.16 T-S Diagram 5.30
5.17 Entropy Change in an Irreversible Process 5.31
5.18 Entropy Change of a Perfect Gas in a Reversible Process 5.33
5.19 Real Heat Engines 5.34
5.20 Principle of Degradation of Energy 5.41
5.21 Theorem of Maximum Work 5.42
5.22 Negative Temperatures 5.43
5.23 Thermodynamic Potentials 5.44
5.24 Maxwell’s Relations 5.48
5.25 How to Remember the Maxwell’s Relations 5.52
5.26 Thermodynamic Variables in Terms of 5.55
Thermodynamic Potentials
5.27 TdS Equations 5.56
5.28 Expressions for Cp – CV 5.57
5.29 Change of Phase: Equilibrium between a Liquid 5.59
and its Vapour
5.30 Clausius-Clapeyron Equation 5.60
Contents xi

5.31 First Order Phase Transition and


Clausius-Clapeyron Equation 5.61
5.32 Second Order Phase Transition 5.64
5.33 The Third Law of Thermodynamics 5.68
5.34 Apparent Violations of the Third Law 5.70
Questions 5.82
Objective Type Questions 5.84
6. Radiation 6.1
6.1 Introduction 6.1
6.2 Heat Waves 6.1
6.3 Detection and Measurement of Thermal Radiation 6.2
6.4 Emission and Absorption of Radiation 6.4
6.5 Prevost’s Theory of Exchanges 6.6
6.6 Emissive Power 6.6
6.7 Radiation in an Enclosure 6.7
6.7 Kirchhoff’s Law 6.8
6.8 Blackbody 6.9
6.9 Intensity and Energy Density 6.10
6.10 Pressure due to Radiation 6.13
6.11 Stefan-Boltzmann Law 6.14
6.12 Verification of Stefan’s Law 6.17
6.13 Determination of Stefan’s Constant in the Laboratory 6.18
6.14 Newton’s Law of Cooling 6.20
6.15 Adiabatic Expansion of Blackbody Radiation 6.20
6.16 Effect of Adiabatic Expansion on Blackbody Radiation 6.21
6.17 Wien’s Law 6.22
6.18 Wien’s Formula 6.25
6.19 Rayleigh-Jeans Formula 6.26
6.20 Planck’s Formula 6.30
6.21 Determination of the Solar Constant 6.33
6.22 Temperature of the Sun 6.34
Questions 6.36
Objective Type Questions 6.37
References R.1
Index I.1
1
Chapter

Basic Ideas

1.1 THERMODYNAMICS
Thermodynamics is the branch of physics in which we study the relationship
between the heat energy and mechanical energy or in general, any other form
of energy. In thermodynamics, we neither consider the internal structure of the
substance, nor we consider the motion of its particles, rather we consider only the
macroscopic properties of the substance.

1.2 ZEROETH LAW OF THERMODYNAMICS AND


TEMPERATURE
According to the Zeroeth Law, if two systems are separately in thermal equilibrium
with a third system then they both are also in thermal equilibrium with each
other.
According to Zeroeth law of thermodynamics, temperature is the quantity
which determines the direction of flow of heat when the two systems are kept
in contact. If there is no exchange of heat energy between the two systems kept
in contact, the systems are said to be in thermal equilibrium (i.e., at the same
temperature). But if there is a transfer of heat energy from one system to the other
system, the system imparting heat energy is said to be at a higher temperature and
the system which receives the heat energy is said to be at a lower temperature.
Temperature is thus the thermodynamic property which controls the state of
thermal equilibrium of the systems.

1.3 TEMPERATURE AND HEAT


Except for judging by the feel of our skin, which is not a very accurate quantitative
guide, the only way we can measure temperature is to measure the effects
temperature changes have on the physical properties of materials e.g., thermal
expansion of certain liquids such as mercury, change in electrical resistance of
a wire, the variation with temperature of electric current generated by dissimilar
metals joined together and so on. The most commonly employed effect to measure
temperature is that of expansion. Thermometers are usually constructed in such a
way that a small thermal expansion results in a large displacement of an indicator.
1.2 Heat and Thermodynamics

In thermometers based on the expansion of a Hot


liquid, a comparatively large amount of liquid
(usually mercury or alcohol) is confined in a
bulb, and its expansion causes the excess liquid
to rise in a narrow capillary tube (Fig. 1.1) shows
a mercury thermometer.
Whatever kind of thermometer we choose,
the scale on which the indicator moves must be
calibrated in some definite easily reproducible
manner. The two thermometric scales in general
use are calibrated at the freezing point and boiling
Cold
point of water (at a pressure of one standard
Hg
atmosphere). On the Centigrade or Celsius
scale, which is used generally all over the world,
Fig. 1.1 A mercury thermometer
the freezing point is 0°C and the boiling point
is 100°C. On the Fahrenheit scale, 32°F is the
freezing point and the boiling point is 212°F (Fig. 1.2). On the Fahrenheit scale,
the interval between freezing and boiling point is divided into 180 degrees (212-
32), and on the Celsius scale, there are 100 degrees in this same interval. The
Fahrenheit and Celsius scales are thus related as
TC TF – 32
___ = _______
5 9
9
e.g., 80°C corresponds to 80 × __ + 32 = 176°F.
5
°F °C
°F °C °F °C
212 100
212° 100°
y 80

32° 0°

68 x

32 0

Crushed ice Boiling


and water water
A B C
Fig. 1.2 Calibration of Fahrenheit and Centigrade (or Celsius) thermometers
Basic Ideas 1.3

1.3.1 Pressure
The pressure P is defined as the magnitude of the force per unit area and the unit
of pressure in the MKS system is 1 newton per square meter (1 Nm–2). A pressure
of exactly 105 Nm–2 (= 106 dyne.cm–2) is called 1 bar, and a pressure of 10–1 Nm–2
(= 1 dyne cm–2) is 1 microbar (1 mbar).
A pressure of 1 standard atmosphere (atm) is defined as the pressure
produced by a vertical column of mercury exactly 76 cm in height, of density
r = 13.5951 g cm–3 at a point where g has its standard value of 980.665 cm s–2.
From the equation P = rgh, we find
1 standard atmosphere = 1.01325 × 106 dyne cm–2
= 1.01325 × 105 Nm–2
Hence, 1 standard atmosphere is very nearly equal to 1 bar, and 1 m bar is very
nearly 10–6 atm.
A unit of pressure commonly used in experimental work at low pressures is
1 Torr (named after Torricelli (1608-1647)) and defined as the pressure produced
by a mercury column exactly 1 mm. in height, under the above conditions.
Therefore
1 Torr = 133.3 Nm –2

1.4 GAS THERMOMETERS


Thermometers using liquid materials (e.g., mercury, alcohol or water) will disagree
among themselves in smaller details due to the fact that different materials react
somewhat differently to an increase of temperature. Thus, we need some other
solution for an exact and universal definition of temperature scale. The solution
is provided by gases since it has been observed that all gases subjected to heating
expand in almost exactly the same way. Thus, we can accept as a standard the
temperature scale provided by a gas thermometer (Fig. 1.3), regardless of what
gas is used to fill it.

1.4.1 Empirical Scale of Temperature


The temperature is defined as a quantity, which prescribes a thermal equilibrium
between two bodies in contact. Among a group of bodies, a particular body A may
be used as a thermometer. The temperature of other body may be compared by
bringing the body A in contact with it. The body used as a thermometer should
posses an easily observable thermometric property, such as the length of a mercury
column in a capillary tube, the pressure of a gas in a bulb, or the resistance of
a platinum wire. The common temperature scale is the Celsius or Centigrade
scale. On this scale, the melting point of ice and boiling point of water at normal
pressure are 0°C and 100°C respectively.
1.4 Heat and Thermodynamics

Open
top

Bulb containing
known volume Mercury
of any gas

Rubber
tube

Fig. 1.3 A gas thermometer. The changing volume of the trapped gas is read from the height of the
mercury column in front of the scale. (In order to keep the mercury level the same on both sides, the right
hand tube is lowered as the gas heats; the gas in the bulb will therefore remain at atmospheric
pressure at all temperature)

Let the temperature t be the linear function of the thermodynamic property X.


If X0, X100 and Xt be the values of X at 0°C, 100°C and t°C respectively, then
X t – X0
t°C = _______ × 100 (1)
X100 – X0
Although the choice of a temperature scale at this point is arbitrary, its
absolute choice will be established with the introduction of the second law of
thermodynamics. The absolute temperature will be represented by T.
We use gas thermometer for an empirical temperature scale. An arbitrary
temperature scale q can the converted into T K by
(PV)qs
T
__ = lim _____ (2)
T0 p Æ 0 (PV)q
0
Basic Ideas 1.5

Thus, T is established uniquely, since T0 is known. If the interval between


steam and ice-points is given a value 100, then
T (PV)qs
__s = lim _____ = 1.366 (3)
T0 P0 Æ 0 (PV)q
0

Ts – T0 = 100 (4)
Equations (3) and (4) define the Kelvin or Absolute Scale. The solution of
these equations gives
T0 = 273.16 K
Ts = 373.16 K
The absolute and centigrade temperatures are therefore related by
T K = 273.16 + t°C (5)

1.5 ABSOLUTE ZERO


Let us take a gas thermometer (Fig. 1.3) and carefully measure the gas volume,
first at the boiling point of water and then again when the thermometer is in ice (In
defining the Celsius temperature scale, we called these temperatures 100°C and
0°C). We can now plot them as points B and F on a graph against our measured
volumes (Fig. 1.4). A straight line through these two points can be drawn and
extended to left and right to define a scale for measurement over a wide range of
temperature.

Fig. 1.4 The behaviour of a gas at low temperature


1.6 Heat and Thermodynamics

There is something peculiar that should be noted about the left (low-temperature)
side of the graph. We cannot extend the line indefinitely as we can on the other
(high-temperature) side, because it soon runs into the axis of the graph indicating
a zero volume. Then it would be ridiculous to extend the line any farther since,
as far as we know, a gas with a negative volume is an idea that has no meaning.
So this point, where our graph seems to be heading, is called the absolute zero of
temperature.
The apparent intention of the gas to shrink to zero volume at absolute zero is
naturally never fulfilled. All gases liquefy before this point is reached; in fact even
before it begins to liquefy, a gas commences to deviate considerably from its more
regular behaviour at higher temperatures. Ingenious experimental procedures and
theoretical corrections, however, have revealed that the temperature of absolute
zero is – 273.16°C. A temperature scale beginning with 0 at absolute zero is an
absolute temperature scale.
The most common absolute scale uses the same size degree as the Celsius
scale and is called the Kelvin scale (K), named after Lord Kelvin (England). Since
the Kelvin degree is the same size as the Celsius degree, and begins counting 273°
lower on the scale, it follows that in order to convert a temperature given in °C
into K, we need only to add 273. This concept of an absolute zero is an important
one in physics, since the triple-point of water is – 273.16°C and internationally
chosen as the standard fixed point for the absolute scale.

1.6 TRIPLE‐POINT CELL


Before 1954, the international metric temperature scale was the Celsius scale,
which was based on the temperature interval between two fixed points:
1. the temperature at which pure ice coexisted in equilibrium with air-
saturated water at standard atmospheric pressure (the ice-point), and
2. the temperature of equilibrium between pure water and pure steam at
standard atmospheric pressure (the steam point).
The temperature interval between these two fixed points was assigned 100
“degrees” (of hotness) i.e., 100°C. Hundreds of attempts were made all over the
world to measure the temperature of the ice point with great accuracy-without
much success. The main difficulty was achieving equilibrium between air-
saturated water and pure ice. When ice melts, it surrounds itself with pure water
that prevents intimate contact between ice and air-saturated water. Attempts to
measure the steam point also presented problem, because the temperature of the
steam point is very sensitive to pressure.
In 1954, a single fixed point was chosen as the basis for a new international
temperature scale, the Kelvin scale. The state in which ice, liquid water and water
Basic Ideas 1.7

vapour coexist in equilibrium (a state known as the triple point of water), provides
the standard reference temperature. The temperature of the triple point of water
which can be very accurately and reproducibly measured was assigned the value
273.16 Kelvin, corresponding to 0.01°C, in order to maintain the magnitude of
a unit of temperature. Note that the word “degree” has been dropped from the
Kelvin scale, so the triple point temperature is abbreviated as 273.16 K. This
temperature is the standard fixed point of thermometry.
To achieve the triple point, water of the highest purity which has substantially
the isotopic composition of ocean water is distilled into a vessel like that shown
schematically in Fig. 1.5.

Fig. 1.5 Triple-point cell with a thermometer in the well, which melts a thin layer of ice nearby

When all air has been removed, the vessel is sealed off. With the aid of a
freezing mixture in the inner well, a layer of ice is formed around the well. When
the freezing mixture is removed and replaced with a thermometer, a thin layer of
ice is melted nearby. So long as the solid, liquid and vapour coexist in equilibrium,
the system is at the triple point.
Celsius temperature t (i.e., centigrade temperature) is defined by the equation
t = T – Ti
where T is the absolute or thermodynamic temperature and Ti is the thermodynamic
temperature of the ice point (= 273.15 K).
At the ice point, T = Ti, t = 0°C and at the triple point of water, T = 273.16 K
t = 0.01°C and at the steam point t = 100°C. Some Kelvin and Celsius temperatures
are compared in Fig. 1.6.
1.8 Heat and Thermodynamics

K C
Steam point 373 K 100°C
100 Kelvins
100 deg C
Ice point 273 K 0°C

NSP CO2 195 K –78°C

NBP oxygen 90 K –183°C

Absolute zero 0 –273°C

Fig. 1.6 Comparison of some Kelvin and Celsius temperatures. Temperatures have been
rounded off to the nearest degree

1.7 TEMPERATURE MEASUREMENT WITH A CONSTANT‐


VOLUME GAS THERMOMETER
The standard thermometer, against
which all other thermometers are
calibrated, is based on the pressure
of a gas in a fixed volume. Figure 1.7
shows such a constant volume gas
thermometer. It consists of a gas-filled
bulb connected by a tube to a mercury
manometer. By raising or lowering
reservoir R, the mercury level on the
left can always be brought to the zero
of the scale to keep the gas volume
constant.
The temperature of anybody in
thermal contact with the bulb is then
defined to be
T = Cp (1)
Fig. 1.7 A constant volume gas thermom-
where p is the pressure within the gas eter. Its bulb is immersed in a liquid whose
temperature T is to be measured
and C is a constant. The pressure
p = p0 – r hg (2)
p0 is the atmospheric pressure, r is the density of mercury in the manometer, h
is the measured difference between the mercury levels in the two arms of the
manometer tube and g is acceleration due to gravity (1 atm = 1.01 × 105 Pa = 760
torr).
Basic Ideas 1.9

If we next put the bulb in a triple point cell, the temperature now being measured
is
T3 = Cp3 (3)
where p3 is the gas pressure now.
Eliminating C between Eqns. (1) and (3), we get temperature
p p
T = T3 __ ( ) __
p3 = (273.16 K) p3 ( ) (4)

We face a problem with this thermometer: if we use it to measure (say) the


boiling point of water, we find that different gases in the bulb give slightly different
results. However, as we use smaller and smaller amounts of gas to fill the bulb,
the readings converge to a single temperature, no matter what gas we use. Thus,
temperature
p
T = (273.16 K) ( lim __
gas Æ 0 p3 ) (5)

and unknown temperature T can be calculated, knowing p and p3.

1.8 INTERNATIONAL PRACTICAL TEMPERATURE SCALE


The accurate measurement of temperature by a standard gas thermometer is a
difficult and tedious process. To help in the calibration and correction of other
thermometers, several basic fixed points have been measured on the Celsius scale
with great accuracy using helium gas constant volume thermometer (Table 1.1).
Between these points, temperatures are easily interpolated by means of resistance
thermometers (– 190°C to 630.5°C), thermocouples (630.5°C to 1063°C) and
radiation thermometers (beyond 1063°C). To facilitate measurements, secondary
fixed points have been added to the International Practical Temperature Scale
based on substances easily available in pure form.

Table 1.1 Basic fixed points on the International Practical Temperature Scale of 1948.
Substance Designation °C °K
Oxygen normal boiling point – 182.97 90.18
Water normal freezing point 0.00 273.15
standard triple point 0.01 273.16
normal boiling point 100.00 373.15
Sulphur normal boiling point 444.60 717.75
Antimony normal melting point 630.5 903.6
Silver normal melting point 960.8 1233.9
Gold normal melting point 1063 1336
1.10 Heat and Thermodynamics

1.9 RESISTANCE THERMOMETER


Pure platinum (Pt) is a suitable metal for the resistance thermometer due to its
resistance to chemical attack and its high melting point (1770°C). A long fine
platinum wire is wound around a thin frame in such a way that excessive strain is
avoided when the wire contracts on cooling.
The resistance of the Pt wire, Rt, in a resistance thermometer with compensating
leads can be measured by a Wheatstone bridge (Fig. 1.8), with RP and RQ as
known fixed resistances and R as an adjustable measuring resistance. We have

Rt = (RQ /RP)R.

Fig. 1.8 Resistance thermometer with compensating leads

The resistance Rt can also be measured by passing a known constant current in


the thermometer and measuring the potential difference across it with the help of
a potentiometer (Fig. 1.9). The current is kept constant by adjusting a rheostat so
that the potential fall across a standard resistance in series with the thermometer,
as checked with a monitoring potentiometer, remains constant.
Basic Ideas 1.11

Fig. 1.9 Potentiometer method measuring Rt

From the oxygen point (– 182.97°C) to the triple point of water (0.01°C), the
temperature t is given by
Rt = R0 [1 + At + Bt2 + C (t – 100) t3]
where Rt is the resistance of the platinum wire at t°C and R0 at 0°C. The constants
R0, A, B and C are determined by measurements at the oxygen point, the triple
point of water, the steam point and the sulphur point.
From the triple point of water to the antimony point (630.5°C), the temperature
t is given by
Rt = R0 [1 + At + Bt2]
The constants are determined by the triple point of water, the steam point and
the sulphur point.

1.10 THERMOCOUPLE
When two wires of different metals are joined to form a closed circuit, an electric
current flows round the circuit so long as the two junctions are kept at different
temperatures. The thermal emf E produced is measured with a potentiometer
(Fig. 1.10). The thermocouple is calibrated by measuring the emf at various
known temperatures, the reference junction being kept at 0°C. The temperature t
at the test junction is given by
E = a + bt + ct2
for the range 650.5°C to the gold point (1063°C).
1.12 Heat and Thermodynamics

Fig. 1.10 Thermocouple

1.11 IDEAL GASES AND GAS LAWS


The gas which perfectly obeys the Boyle’s and Charle’s laws is called an ideal or
perfect gas.

Assumptions for an Ideal Gas


(i) This gas obeys perfectly the Boyle’s, Charle’s and pressure laws at
each pressure and temperature.
(ii) The coefficient of volume expansion and the pressure coefficient of
this gas are equal = 1/273 °C–1
(iii) The molecules are of negligible size in comparison to the volume of
the gas.
(iv) There are no forces of attraction amongst the molecules of the gas. Thus
an ideal gas cannot be obtained in the liquid or solid state. Indeed, no
gas is a perfectly an ideal gas, but in practice we can assume oxygen,
nitrogen, hydrogen and helium to be nearly the ideal gases at ordinary
temperature and ordinary pressure because it is difficult to liquefy
them under these conditions.
Boyle’s Law At a constant temperature, the volume of a given mass of gas is
inversely proportional to its pressure:
1
V μ __
P
or PV = constant (at T = constant)
or P1V1 = P2V2 (1)
Basic Ideas 1.13

If a graph is plotted by taking the pressure P on


the Y axis and volume V on the x-axis, a rectangular
hyperbola is obtained (Fig. 1.11). This curve is called
an isothermal curve.
Charle’s Law At a constant pressure, the volume of a

P
given mass of a gas increases (decreases) by (1/273)
of its volume at 0°C on increasing (decreasing) the
temperature by 1°C. V
Let at a constant pressure, the volume of given mass Fig. 1.11 P-V graph
of a gas at 0°C be V0 and at t°C be Vt. Since increase in
volume of gas on increasing its temperature by t°C will be V0t/273, therefore, the
volume of a gas at t°C is given by
V0t
Vt = V0 + ____
273
t
or [
Vt = V0 1 + ____
273 ] (2)

If a graph is plotted, taking volume V on Y-axis and temperature t on X-axis,


we get a straight line (Fig. 1.12) which meets the temperature axis at – 273°C.
Volume V

Vt

V0

– 273 0 t
Temperature t (in °C)

Fig. 1.12 V-t graph

Charle’s Law in Terms of Absolute Temperature


Let, at a constant pressure, the volume of given mass of a gas at 0°C, t1°C and t2°C
be respectively V0, V1 and V2, then by Charle’s law.
t1
(
V1 = V0 1 + ____ ;
273 )
t2
(
V2 = V0 1 + ____
273 )
1.14 Heat and Thermodynamics

V1 [1 + (t1/273)]
__ = ___________
V2 [1 + (t2/273)]
273 + t1 T1
= _______ = __
273 + t2 T2
V
__
or = constant (3)
T
i.e., V μT
Thus, at a constant pressure, the volume of given mass of a gas is directly
proportional to its absolute temperature.

Pressure Law
Keeping the volume constant, if given mass of gas is heated, its pressure increases.
Experimentally, it is found that at a constant volume, the pressure of a given
mass of a gas increases (decreases) by 1/273 of its pressure at 0°C or increasing
(decreasing) its temperature by 1°C.
Let at a constant volume, the pressure of given mass of a gas at 0°C be P0 and
at t°C be Pt. Then by pressure law

P0t
Pt = P0 + ____
273
t
or [
Pt = P0 1 + ____
273 ] (4)

If a graph is plotted by taking the pressure P on Y-axis and temperature t on


X-axis, a straight line is obtained (Fig. 1.13) which meets the temperature axis
at – 273°C.
Pressure P

Pt

P0

– 273 0 t
Temperature t (in °C)

Fig. 1.13
Basic Ideas 1.15

Pressure Law in Terms of Absolute Temperature


Let at a constant volume, the pressure of given mass of a gas at 0°C, t1°C and t2°C
be respectively P0, P1 and P2, then by pressure law

t1
(
P1 = P0 1 + ____ ;
273 )
t2
(
P2 = P0 1 + ____
273 )
P1 1 + (t1/273)
\ __ = __________
P2 1 + (t2/273)

273 + t1 T1
= _______ = __
273 + t2 T2

P
__
or = constant
T
i.e., PμT (5)
Thus, at a constant volume, the pressure of a given mass of a gas is directly
proportional to its absolute temperature.

1.12 AVOGADRO’S NUMBER


While thinking of molecular aspect of gases, one has to measure the sizes of gas
– samples in moles. If we do so, then we can be certain that we are comparing
samples that contain the same number of atoms or molecules. The mole is one of
the seven SI base units and is defined as follows:
One mole is the number of atoms in a 12g sample of Carbon-12.
Then one may enquire “How many atoms or molecules are there in a mole”?
The answer is determined experimentally and is

NA = 6.02 × 1023 mol–1 (1)


(This is known as Avogadro’s number).
Here mol–1 represents “per mole”. This is based on Avogadro’s assertion that
all gases contain the same number of atoms or molecules when they occupy the
same volume under the same conditions of temperature and pressure.
The number of moles n contained in a sample of any substance having N
molecules is given by
N
n = ___ (2)
NA
1.16 Heat and Thermodynamics

1.13 UNIVERSAL GAS CONSTANT, BOLTZMANN CONSTANT


AND IDEAL GAS EQUATION
Experiments have confirmed that if we confine 1 mole samples of various gases in
boxes of identical volume and hold the gases at the same temperature, then their
measured pressures are nearly (though not exactly) the same. If the measurements
are at lower gas densities, then the small differences in the measured pressures
tend to disappear. Further, at low enough densities, all real gases tend to obey the
relation.
pV = nRT (ideal gas law) (1)
where p is the absolute (not gauge) pressure, n is the number of moles of gas
present and T is the absolute temperature (in Kelvin). The symbol R is a constant
called the Universal Gas Constant (that has the same value for all gases).
R = 8.31 J/mol.K (2)
Equation (1) is called the ideal gas law. Provided the gas density is low, this
law holds for any single gas or any mixture of different gases (For a mixture, n is
the total number of moles in the mixture).
We can rewrite Eqn. (1) in an alternative form in terms of a constant called the
Boltzmann constant, which is defined as

R 8.31 J/mol.K
k = ___ = _______________
NA 6.02 × 1023 mol–1

= 1.38 × 10–23 J/K (3)


\ R = kNA
N N
But n = ___ or NA = __
n (4)
NA
kN
\ R = ___
n (5)

or nR = Nk (6)
Substituting this into Eqn. (1) gives a second expression for the ideal gas law:

pV = NkT (ideal gas law) (7)

(Note the difference between the Eqns. (1) and (7), for the ideal gas law: Eqn.
(1) involves the number of moles n, whereas Eqn. (7) involves the number of
molecules N).
The data for CO2 are plotted in Fig. 1.14 for three different temperatures. The
remarkable feature of these curves is (a) that they all converge to exactly the same
point on the vertical axis, whatever the temperature and (b) that the curves for all
Basic Ideas 1.17

other gases converge to exactly the same point. This common limit of the ratio
Pv/T as P approaches zero is, the universal gas constant (R). The unit of Pv/T is

1 (Nm–2) (m3 kilomole–1) (K–1) = 1 (Nm) (Kilomole–1 K–1)


= 1 J kilomole–1 K–1
and the value of R in this system is
R = 8.3143 × 103 J kilomole–1 K–1
At sufficiently low pressures, we can write for all gases
Pv/T = R or Pv = RT

Pv
___
or = R (ideal gas)
T
Since v = V/n
\ PV = nRT
For an ideal gas, the curves in Fig. 1.14 coalesce into a single horizontal straight
line at a height R above the pressure axis.

10
3
R = 8.3143 × 10
)

Ideal gas
–1

8
T3
Pv/(J kilomole K
–1

6
T2
4
T3 > T2 > T1
2 T1

7
0 2 4 6 8 × 10
–2
Pressure (Nm )

Fig. 1.14 The limiting value of Pv/T is independent of T for all gases. For an ideal gas,
Pv/T is constant

1.14 THE “SYSTEM”


In order to understand the world around us, it is just not possible to make
observations on the whole world at a time. For our convenience, we can extract
relevant portions of the world (depending on the view to use it to our advantage)
and put them under observations. These relevant parts of the physical world
separated from the rest of the world by means of well-defined boundaries are
called systems. Whatever is left out is called environment or surroundings.
1.18 Heat and Thermodynamics

If ‘mass’ is held constant the system is called a closed system. The surrounding
walls of a closed system must not permit any material transport across them.
Although mass of a closed system remains constant, yet the energy may vary. This
implies that surrounding walls though impermeable to matter should be permeable
to energy (This implies that the walls should be made of good conductors of heat
or electricity).
If the enclosing walls are impermeable to both matter and energy, the system
is said to be an isolated system. Thus, the enclosing walls of an isolated system
should be impervious, rigid and made of non-conducting materials (One of the
nearest approaches to an isolated system is a Dewar flask).
The mass and energy of an isolated system remains constant because it does not
interact with its environment in terms of both mass transfer and energy transfer.
Thus, the isolated system though highly idealized and practically unattainable is
a useful concept in performing energy analysis because system plus environment
can always be taken as an isolated system.
If a system communicates with its environment in terms of both mass transfer
and energy transfer, it is called an open system.
Apart from the terms ‘closed’, ‘open’ and ‘isolated’, there are also other
adjectives to qualify systems further. These are, for example, isothermal (meaning
same temperature), isobaric (meaning same pressure), adiabatic (meaning
no exchange of heat with the environment), isochoric (meaning no change of
volume) etc. These adjectives can be used to qualify a process e.g., an isothermal
process (i.e., a process at constant temperature) isochoric process (a process in
which volume remains constant).
Further, the minimum number of independent variables necessary for
complete description of the system are called state variables. These variables
define the state of the system. State variables in fact are mostly the experimentally
determinable properties of a system (Property is an attribute which belongs to a
system and which, in principle, can be quantitatively evaluated). It must be borne
in mind that property should be relevant to the context. For example, while talking
about thermodynamics, the relevant properties are energy, volume, temperature,
pressure, etc. Note that heat and work are not properties as they do not belong to
the system. These are operations which are performed on the system to alter its
energy. In thermodynamics, the central theme is energy and energy transfer. The
relevant properties are therefore, those that are involved with energy in some way
or the other.

1.15 INTENSIVE AND EXTENSIVE PROPERTIES


Imagine a system in equilibrium to be divided into two equal parts, each with
equal mass. Those properties of each half of the system that remain the same to be
Basic Ideas 1.19

intensive and those that are halved are called extensive. The intensive coordinates
of a system, such as temperature and pressure, are independent of the mass; the
extensive properties are proportional to the mass. Extensive properties are thus
additive i.e., their value for the whole system is equal to the sum of the values for
the individual parts. Intensive properties are not additive.
An extensive variable X, when divided by the mass m or the number of moles
n of system, becomes an intensive variable x
X X
x = __
m or x = __
n
nM = m

where M is the molecular mass (or molecular weight).


The ratio of an extensive variable to the mass of the system is called the specific
value of that variables e.g., the volume V and specific volume (or volume per unit
mass) v are related by
V
v = __
m
The ratio of an extensive variable to the number of moles of a system is called
the molar (or molal) specific value of that variables e.g., molar specific volume
v, is
V
v = __
n
No confusion arises from the use of the same letter to denote both the extensive
variable per unit mass and the extensive variable per mole. In almost every
equation in such a variable occurs, there comes some other quantity that indicates
which variable, X/m or X/n, is meant. If in some rare case there happens to be no
such indication, the equation holds equally well for either case.

1.16 THERMODYNAMIC SYSTEM AND THERMODYNAMIC


COORDINATES
A system whose state depends on the thermodynamic quantities is called
thermodynamic system. For example a gas enclosed in cylinder is a thermodynamic
system because state of the gas depends on its pressure, volume and temperature.
A thermodynamic system means a macroscopic body in equilibrium which is
made up of large number of microparticles e.g., molecules, atoms or ions etc.
The quantities required for the complete knowledge of thermodynamic state
of a system are called the thermodynamic coordinates e.g., pressure P, volume
V and temperature T are the thermodynamic coordinates of the gaseous systems.
These quantities are related by an equation called the equation of state of the
1.20 Heat and Thermodynamics

system (e.g., PV = RT is the equation of state for a perfect gas). It is clear from
the equation of state that out of the thermodynamic coordinates P, V and T, only
two coordinates are independent variables, which can be chosen arbitrarily, the
third coordinate is the dependent variable and can be determined with the help of
equation of state.

1.17 P‐v‐T SURFACE FOR AN IDEAL GAS AND BOYLE’S LAW


The equation of state of a PvT system defines a surface in a rectangular coordinate
system in which P, v, and T are plotted along the three axes. A portion of this
surface for an ideal gas is shown in Fig. 1.15. Every possible equilibrium state
of an ideal gas is represented by a point on its P-v-T surface, and every point
on the surface represents a possible equilibrium state. A quasistatic process, i.e.,
a succession of equilibrium states, is represented by a line on the surface. The
full lines in Fig. 1.15 represent processes at constant temperature, or isothermal
processes. The dotted lines represent isochoric processes, and the dashed lines,
isobaric processes.
Pressure

Vo
lume e
tur
pera
Tem

Fig. 1.15 P-v-T surface for an ideal gas


Basic Ideas 1.21

Figures 1.16(a) and 1.16(b) are projections of the lines in Fig. 1.15 onto the P-v
and P-T planes.

Pressure
Pressure

Volume Temperature

(a) (b)

Fig. 1.16 Projections of the ideal gas P-v-T surface onto (a) the P-v plane, and (b) the P-T plane

In an isothermal process, for a fixed mass of an ideal gas,

Pv = RT = constant (1)

Robert Boyle, in 1660, discovered experimentally that the product of the


pressure and volume is very nearly constant for a fixed mass of a real gas at
constant temperature. This fact is known as Boyle’s law. It is, of course, exactly
true for an ideal gas, by definition. The curves in Fig. 1.16(a) are graphs of Eqn. (1)
for different temperatures and hence for different values of the constant. They are
equilateral hyperbolas.
In a process at constant volume, for a fixed mass of an ideal gas,

( )
nR
P = ___ T = constant × T
V
(2)

That is, the pressure is a linear function of the temperature T. The dotted lines
in Fig. 1.16(b) are graphs of Eqn. (2) for different volumes and hence different
values of the constant.
If the pressure of a fixed mass of an ideal gas is constant,

( )
nR
V = ___ T = constant × T
P
(3)

and the volume is a linear function of the temperature at constant pressure.


1.22 Heat and Thermodynamics

1.18 BEHAVIOUR OF REAL GASES EQUATION OF STATE


AND P‐v‐T SURFACES
The assumption of non-interacting molecules is satisfactory upto a point. After all,
the elastic collision between two molecules implies a repulsive force at sufficiently
small distances. Figure 1.17(a) is a sketch of potentials of such a force where ro
represents the radius of the hard-sphere model molecule. A mutual interaction
between molecules of a real gas comes into existence due to rearrangement of the
charge distribution in each of the molecules.

van der Waal’s Interaction


The interaction is explained in terms of van der Waals forces, as follows:
Consider two nearby atoms. It is true that on an average, the centre of negative
charge in an atom coincides with the centre of positive charge (i.e., with the
nucleus). But the electrons are moving around the nucleus and at any instant it may
happen that the two centres fail to coincide and an instantaneous electric dipole
moment of magnitude m1 develops which produces an electric field E = 2m1/r3
at the centre of the second atom distant r from the first atom. This field induces
an instantaneous dipole moment m2 = aE = 2am1/r3 on the second atom where
a is the electronic polarizability. The interaction potential energy of the dipole
moments is
Ve(r) = – m2E = – aE2
4am12 c
= – _____ ∫ – __6 (1)
r6 r
This is called van der Waals interaction.
The negative sign shows that the interaction is attractive (Ve ~ – 2 × 10–14
erg).
As the atoms (or molecules) approach each other more closely, a repulsive
term arises from the action of electric charges of one atom on the other (This
repulsive force must exist otherwise two atoms would not repel one another after
impact). Experimental data on the inert gases can be fitted well by assuming that
the repulsive potential is of the form B/r12, where B is a positive constant, when
used with a long range attractive potential of the form (Eqn. 1). Then, we can
write the total potential energy of two interacting atoms at separation r as

V(r) = 4 e [ ( __sr ) – ( __sr ) ]


12 6
(2)

where e and s are the parameters with


4 e s6 = C
and 4 e s12 = B (3)
Basic Ideas 1.23

The potential (2) is called the Lennard-Jones potential and is plotted in


Fig. 1.17(a). The minimum occurs at ro = s 21/6.

Potential Energy V
Potential Energy V (r)

Repulsion
Separation Separation
r
s e
e

d
r0
Attraction
(a) (b)

Fig. 1.17 (a) Lennard-Jones potential (b) Idealized interatomic potential

The value of V(r) at the minimum is – e and V(r) = 0 at r = s. The constants B


and C (or e and s) are empirical parameters determines from measurements made
on gases (The data used include the virial coefficients and the viscosity). The
point of minimum V(r), ro is the equilibrium separation at 0°K and corresponds
to the density of the solid. Gases always have r > ro. van der Waals idealized the
situation by approximating the repulsive part by an infinite hard sphere repulsion
so that the potential energy looks like that shown in Fig. 1.17(b). Thus, each atom
is imagined to be a hard sphere of radius ro surrounded by an attractive force
field. The effects of the repulsive and attractive parts are then taken into account
separately. Some typical values of ro and e are given in Table 1.2.

Table 1.2 Values of ro and e.


Molecule 2ro (Å) e (ergs)
He 2.2 1 × 10–15
H2 2.7 4 × 10–15
Ar 3.2 5 × 10–15
CO2 4.5 40 × 10–15

The van der Waals force (F = – ∂V/∂r) is repulsive for distances less than
~ 2.5 × 10–10 m which is of the order of diameter of the molecule. The attractive
force becomes vanishingly small for distances ~ four times the diameter. Within
1.24 Heat and Thermodynamics

this range the attractive force is about 5 × 10–13 Newtons. Though this force is
small, considering the fact that molecular masses are of the order of a 10–26 kg,
this attractive force plays a significant role in molecular motion.

Modification of the Ideal Gas Equation


The above fact suggests that the ideal gas equation
PV = nRT (4)
needs to be modified for the case of a real gas. The need for such a modification is
more so if the above equation is to be employed in the measurement of temperatures
or calibration of standard thermometers employing real gases. The modifications
as applicable to real gases on the following lines were first suggested by van der
Waals.
As a consequence of the finite size of a molecule, the centre of the molecule
cannot approach the wall of the container closes than ro. The minimum distance
between two colliding molecules is 2ro. Hence the actual volume available for the
molecule is less than the volume of the container. The precise decrement depends
on the number of molecules. Let us write the gas equation for one mole of gas i.e.,
in terms of the molar volume v (= V/n). If v is the molar volume with a real gas,
to achieve the same pressure with point-mass molecules of an ideal gas we would
need a volume less than v by an amount b called the covolume. Therefore,
P (v – b) = RT (5)
The effect of short range forces amongst the molecules is to render the
molecular density somewhat more compact on the macroscopic scale. The internal
pressure in the gas is due to the mutual interaction between the molecules. If
there are No molecules per unit volume, the number of mutually interacting pairs
is No (No – 1)/2, which for large N is proportional to No2. Thus, in addition to
the external pressure, there is an additional term which is proportional to N2 or
equivalently to the inverse square of molar volume (i.e., 1/v2). The equation of
state, therefore, is

(P + a/v2) (v – b) = RT (6)

where a is a constant. This is known as the van der Waals equation of state for a
real gas. The term (a/v2) is called internal pressure. It is clear that at low pressure
v >> b and also a/v2 << P, so that Eqn. (6) will closely approximate to the ideal
gas equation. The constants a and b vary from gas to gas. Table 1.3 gives the
values of constants a and b for a few gases.
Basic Ideas 1.25

Table 1.3 Values of van der Waals constants, for P in atm., v in cc/mole, T in K and R = 0.0820575
li atm/K mole.

Substance a b
li2 atm/mole2 cc/mole (li/mole)
He 0.0341 23.7 (0.0237)
H2 0.244 26.6 (0.0266)
CO2 3.59 42.7 (0.0427)
H2O 5.46 30.35 (0.0305)

Figure 1.18(a) is a diagram of a portion of the P-v-T surface of van der Waals
gas and Fig. 1.18(b) is a projection of a number of isotherms onto the P-v plane.

Fig. 1.18 (a) P-v-T surface for a van der Waals gas; (b) Isotherms of a van der Waals gas

Similar sorts of P-v-T surfaces and isotherms can be constructed for other real
substances.

Theoretical P-V Curves (Isotherms) for CO2


From Eqn. (6), we have

RT a
P = _____ – __2 (7)
v–b v
Substituting the values of a and b for CO2 in the above equation, if we plot
a graph of P against v at different constant temperatures, we get the isotherms
shown in Fig. 1.19.
1.26 Heat and Thermodynamics

Above the critical temperature


31.4°C, the theoretical curve appears
to be identical to the Andrew’s
experimental curve, but below the
critical temperature 31.4°C, the main
difference in the theoretical and the
experimental curves is that where
there is maxima and minima in the
theoretical curves, there is a horizontal
line (e.g., CB) in the experimental
curves. For example, at 13.1°C, there Fig. 1.19 P-v isotherms from van der Waals
equation of state for carbon dioxide
is a continuous part BbcC in the
theoretical curve with maxima at b and
minima at c, whereas there is a horizontal line BC in the experimental curve. This
dissimilarity can be explained as follows:
The part bc of the theoretical curve shows that the volume decreases with
the decrease in pressure. This is the unstable state which cannot be achieved in
practice. That is why this state is not obtained in the experimental curve. Thus the
parts Bb and cC of the theoretical curve represent respectively the super-saturated
vapour state and the super-saturated liquid state, which are impossible to obtain
experimentally under ordinary conditions.

1.19 HEAT CAPACITY


Heat capacity of a substance is the ratio of heat supplied to it to the rise in its
temperature. A little consideration would show that this definition would be quite
misleading. Consider for instance, a definite amount of gas at room temperature,
contained in a cylinder fitted with a movable piston. If the piston is pushed in
suddenly the temperature of the gas would rise. The ratio of heat supplied to the
rise in temperature (heat capacity) in this situation is zero.
Let us now perform another experiment. Now, suddenly pull the piston out.
As a result, temperature of the gas will fall. After the temperature has fallen
below the room temperature. Let us start heating the gas till it comes back to its
original temperature. In this experiment, the heat supplied to the system is finite
though the rise in temperature is zero. In this situation, therefore, the ratio of heat
supplied to the rise in temperature would be infinite. Thus, the heat capacity of
the gas can have any value between zero and infinity depending upon how the
experiment for its evaluation has been performed. Such a situation is inexplicable
and it becomes necessary to talk of heat capacity under restrictive conditions. The
restrictive conditions are constant pressure and constant volume. The heat capacity
at constant pressure and the heat capacity at constant volume are represented as
Cp and Cv respectively.
Basic Ideas 1.27

It is well known that Cp is generally greater than Cv. The more rigorous
thermodynamic approach to Cp and Cv will be presented in the next chapter. It
will be seen that in the case of incompressible substances, Cp and Cv are more or
less equal. In the differential form, we can express
(dQ)p
dT
∂Q
∂T( )
Cp = _____ ∫ ___
p

(dQ)v
and
dT ( )
∂Q
Cv = _____ ∫ ___
∂T v

(dQ)p and (dQ)v are heat supplied at constant pressure and constant volume
respectively.

1.20 HEAT: ITS NATURE AND UNITS


Prior to Joule’s work on mechanical equivalent of heat, for many years, the quantity
of heat flowing into a system was expressed in terms of calories or British thermal
units (Btu). 1 calorie was being defined as the heat flow into 1 gm of water in a
process in which its temperature increased by 1 Celsius degree and 1 Btu is the heat
flow into 1 pound – mass of water when its temperature increased by 1 Fahrenheit
degree. Careful measurements showed that these quantities of heat varied slightly
with the particular location of the one degree interval, for example, whether it
was from 0°C to 1°C or from 50°C to 51°C. To avoid confusion, the 15-degree
calorie was defined as the heat flow into 1 gm of water when its temperature was
increased from 14.5°C to 15.5°C.
If the same rise in temperature is produced by the performance of dissipative
work, the best experimental results show that 4.1858 jouls are required, a value
that is referred to as the mechanical equivalent of heat (see next article). We can
then say

One 15-degree calorie = 4.1858 joules.

This relation between the joule and the 15-degree calorie is necessarily subject
to some experimental uncertainty. For this reason, and also so as not to base the
definition of the calorie on the properties of some particular material (i.e., water)
an international commission has agreed to define the New International Steam
Table calorie (the IT calorie) by the equation.
1 3600
One IT calorie = ____ watt hour = _____ joules (exactly)
860 860
Then, to five significant figures
1 IT calorie = 4.1860 joules.
1.28 Heat and Thermodynamics

Here, the figure of 860 was chosen so that the IT colorie chosen would agree
closely with the experimental value of the 15-degree calorie.
Since the relations between the joule and the foot-pound (FPS unit of energy),
between the gram and the pound-mass, and between the Celsius and Fahrenheit
degrees are also matters of definition and not subject to experimental uncertainty.
The British thermal unit is also defined exactly in terms of the joule. To five
significant figures
1 Btu = 778.28 foot-pounds.
For many years, it was thought that heat was a substance contained in material.
The first conclusive evidence that it was not was given by Count Rumford (1753
– 1814) (He observed the temperature rise of the chips produced while boring
cannons. He concluded that heat flow into the chips was caused by the work of
boring). However, the precision measurements of the mechanical equivalent of
heat were made by Joule (next article). The experiments were performed in a
period from 1840 to 1878, and although, Joule expressed his results in English
units, they are equivalent to the remarkably precise value of
1 calorie = 4.19 joules.
(The energy unit, 1 joule, was not introduced or named until after Joule’s death
and the standardized 15-degree calorie had not been agreed on at the time of
Joule’s work).
However, the true significance of Joule’s work went for beyond a mere
determination of the mechanical equivalent of heat. By means of such above
mentioned experiments, and others of a similar nature, Joule concluded that there
was in fact a direct proportion between “work” and “heat” and he succeeded in
dispelling the belief, prevalent at that time that “heat” was an invisible, weight
less fluid known as “caloric”. It may be said that Joule not only determined the
value of the mechanical equivalent of heat but also provided the experimental
proof that such an equivalence actually existed.

1.21 MECHANICAL EQUIVALENT OF HEAT


It is a well known fact that friction produces heat. Friction forces also gradually
rob mechanical systems of their energy and eventually bring them to rest. Then
one may ask what is the relation between the mechanical energy lost to friction
and the amount of heat produced by it. This question was answered by James P.
Joule (1818-1889) in his famous experiment on the transformation of mechanical
energy into heat.
Joule’s apparatus, schematically shown in Fig. 1.20, consisted of a water-filled
vessel containing a rotating axle with several stirring paddles attached to it. The
water in the vessel was prevented from rotating along with the paddles by special
Basic Ideas 1.29

Fig. 1.20 Joule’s apparatus for determining the mechanical equivalent of heat

vanes attached to the walls of the vessel. The axle with the paddles was driven
by a weight hanging from a cord, and thus, the work done by the descending
weight was transformed by friction into heat produced in the water. Knowing the
amount of water in the vessel, Joule could measure the rise of its temperature and
calculate the total amount of heat produced, the driving weight and the distance
of its decent gave the total amount of mechanical work done. Repeating this
experiment many times and under different conditions, Joule established that
there is a direct proportionality between these two qualities and that “the work
done by the weight of 1 pound through 772 feet at Manchester will, if spent in
producing heat by friction in water, raise the temperature of 1 pound of water
one degree Fahrenheit”. In CGS units, it means that one calorie of heat is the
equivalent of 4.18 × 107 ergs of work or 4.18 Joules.
Joules work confirmed the basic ideal that was commencing to be seriously
considered at the time, namely, that heat is energy in the same sense that
mechanical energy is, and although one form of energy can be transformed into
another (kinetic, potential, heat, electrical, chemical, etc.), the total sum of the
energy in any system remains constant. This law of conservation of energy led to
the statement of the first law of thermodynamics.
1.30 Heat and Thermodynamics

1.22 THERMAL EXPANSION


When the temperature of a solid is increased it expands because of the increase in
average distance between the atoms. The change in any linear dimension of the
solid is called linear expansion.
If the change in temperature DT causes a corresponding change Dl in length
from its original length l then the coefficient of linear expansion a is defined as
1 Dl
a = __ ___ (1)
l DT
For isotropic solid, the percent change in length for a given temperature change
is the same for all times. The fraction change in the area A per degree temperature
change is b.
Then,
DA
___ = b DT
A
or DA = bA DT (2)
The fractional change in volume V per degree temperature change is g and is
given by
DV = gV DT (3)
For fluids with no definite shape, particularly the gas, only the change in
volume is significant.
The coefficient of volume expansion is then given by
1 DV
g = __ ___ (4)
V DT
For liquids, the expansion in volume with increasing temperature is about ten
times greater than that of solids.
Example 1.1 Find what value would be assigned to the temperature of boiling
water if one used a CO2 constant volume thermometer (v = 500 cm3 g–1), assuming
that CO2 obeys the van der Waals equation ( p in atmospheres, v in cm3 g–1)
1.8654 T 1861
p = ________ – _____
v – 0.973 v2
Solution On using the thermometer, one would measure the pressure p3 at the
triple point of water and the pressure pb at the boiling point and use the definition
of gas temperature
p
T = 273.16 lim __
pÆ0 p 3( ) (i)

pb
\ Tbp = __
p3 × 273.16 (ii)
Basic Ideas 1.31

Now, from the given formula for p, the measured pressure would be
p3 (T = 273.16) = 1.0136
pb (T = 373.15) = 1.3874
\ From (ii), we get
Tbp = 373.88°K.
Example 1.2 Prove that the change in period t of a physical pendulum with
temperature T is given by
1
Dt = __ at DT
2
Solution The time period
___
t = 2p ÷l/g
1 l–1
\ Dt = 2p × __ _______
__ Dl
2 2/÷g
__
Dt __
__ 1 Dl ÷g __1 Dl
or = ___
__ × ___
_ = __
t 2 gl ÷l 2 l
÷
1 1
= __ a l DT = __ aDT,
2l 2
1
\ Dt = __ at DT.
2
Example 1.3 Density is mass per unit volume. If the volume V is temperature
dependent, so is the density. Show that change in density Dr with change in
temperature DT is
Dr = – grDT
where g is the volume coefficient of expansion.
M
Solution Density r = __
V
M
Dr = – ___2 DV
V
But DV = gV DT

M
Therefore Dr = – ___2 (gV DT)
V
M
= – __ g DT
V
1.32 Heat and Thermodynamics

or Dr = – gr DT
The minus sign implies that r decreases with rise in temperature.
Example 1.4 Show that when the temperature of a liquid in a barometer changes
by DT and the pressure is constant, the height h changes by Dh = gh DT where g is
the coefficient of volume expansion.
Solution The pressure p of the liquid is p = hrg where h is the height of mercury
and r is the density and g, the acceleration due to gravity.
Differentiating, assuming p to be an exact differential, we have
∂p ∂p
dp = ___ Dh + ___ Dr
∂h ∂r
Since p is constant \ dp = 0
0 = rg Dh + hg Dr
Dividing by hrg, we have
Dh Dr
0 = ___ + ___
h r
and using the result of the previous problem
Dr = – gr DT
we get
Dr
Dh = – ___ . h
r
= gh DT
Example 1.5 Show that the volume thermal expansion coefficient for an ideal gas
at constant pressure is 1/T.
Solution Let the initial volume of the gas be Vo at temperature To, at constant
pressure p and DT rise in temperature corresponds to the increase in volume by
DV.
Then, from gas equation pV = nRT, we have
p (Vo + DV) = nR (To + DT) (i)
and pVo = nRTo (ii)
Subtracting,
pDV = nR DT (iii)
nR __ Vo
But ___
p To (from Eqn. (ii))
=
Basic Ideas 1.33

Therefore from (iii)


nR V
DV = ___ __o
p . DT = To DT
DV
or ___
Vo ( )
1
= __ DT
To
DV
___
Since = gDT
Vo
1 1
where g is volume expansion coefficient we find g = __ or g = __
To T
(\ To is arbitrary).
Example 1.6 A body of mass 2 kg is being dragged with a uniform velocity of 2
ms–1 on a rough horizontal plane. The coefficient of friction between the body and
the surface is 0.2. Calculate the amount of heat generated in 5 sec.
J = 4.2 J cal–1, g = 9.8 ms–2.
Solution Force = coefficient of friction × normal reaction
= 0.2 × 2 × 9.8 Newtons
Distance travelled = velocity × time
= 2 × 5 = 10 m
Work done = force × distance
= 0.2 × 2 × 9.8 × 10 J
W 0.2 × 2 × 9.8 × 10
\ Heat produced = __ = _______________ cal
J 4.2
= 9.33 cal.
Example 1.7 The temperature of 25 g of water vapour increases by 9.69 K when
0.45 kJ of heat is supplied at constant pressure. What is its molar constant pressure
heat capacity (Cp), if it behaves as a perfect gas.
Solution Since dQp = nCp dT
25 g
\ 0.45 × 103 J = ________ × Cp × 9.69 K
18 g/mol
Consequently
0.45 × 103 J × 18 mol–1
Cp = ___________________
25 × 9.69 K

= 33.44 J K–1 mol–1


1.34 Heat and Thermodynamics

Example 1.8 Assuming the Avogadro number NA = 6.02 × 1023 mol–1, universal
gas constant R = 8.31 J mol–1 K–1 and the molecular weight of hydrogen = 2,
calculate (i) mass of one molecule of hydrogen (i.e., H2) and (ii) Boltzmann
constant.
Solution
(i) Mass of 1 molecule of hydrogen

Molecular Weight
= ________________
Avogadro’s number
2
= _________ gm = 3.32 × 10–24 gm
6.02 × 1023

= 3.32 × 10–27 kg
(ii) Boltzmann constant

R 8.31
k = ___ = _________
NA 6.02 × 1023

= 1.38 × 10–23 J mole–1 K–1.

Example 1.9 The volume of a gas at a pressure 1.2 × 107 Nm–2 and temperature
400 K is 1 litre. Assuming the Avogadro’s number NA = 6.02 × 1023 per mole,
calculate the number of molecules in the gas.
Solution We know that number of molecules is 22.4 litre of any gas at N.T.P.,
(P = 1.01 × 105 Nm–2 and T = 273 K) = Avogadro’s number NA (= 6.02 × 1023).
Thus, we first need to find the volume of the given gas at NTP

P1 = 1.2 × 107 Nm–2, T1 = 400 K, V1 = 1 litre


P2 = 1.01 × 105 Nm–2, T2 = 273 K, V2 = ?

From perfect gas equation PV/T = R, we have

P1V1 P2V2
____ = ____
T1 T2
P 1 T2
or V2 = __ × __ × V1
P2 T1

1.2 × 107 273


or V2 = _________5 × ____ × 1
1.01 × 10 400

= 81.09 litre
Basic Ideas 1.35

Since at NTP, the number of molecules in 22.4 litre gas is 6.02 × 1023, \
number of molecules in 81.09 litre gas
81.09
= _____ × 6.02 × 1023
22.4
= 2.18 × 1024
Thus at a given temperature and pressure, 1 litre gas will have 2.18 × 1024
molecules.
Example 1.10 Centigrade and Fahrenheit thermometer are immersed in a fluid in
which the Fahrenheit reading is thrice that of the centigrade reading. What are the
absolute values of temperature in °K and °R.
Solution Let the reading of the centigrade thermometer be t°C. Then that in
Fahrenheit thermometer will be 3t°F.
C ______
__ F – 32
Now =
5 9
3t – 32
t__ ______
\ =
5 9
or 6t = 160
\ t°C = 26.7°C
3t°F = 80°F
Hence, the absolute temperature will be
T °K = 273.169 + t°C = 273.16 + 26.7
= 299.86°K
and T °R = 459.8 + 3t°F (\ 0°F Æ 459.8°R)
= 459.8 + 80 = 539.8°R
Example 1.11 The resistances of a platinum resistance thermometer at 0°C
and 100°C are 2.56 and 3.56 ohms respectively. Calculate the true temperature
when the resistance of the thermometer is 5.56 ohms. Given, for pure platinum
a = 3.94 × 10–3, b = – 5.80 × 10–7.
Solution If Rt and Ro be the resistances at t°C and 0°C respectively, then

Rt = Ro (1 + at + bt2) (i)
where a and b are constants. Then, the temperature tpt measured by Pt thermometer
is
Rt – Ro
tpt = _______ × 100 (ii)
R100 – Ro
1.36 Heat and Thermodynamics

where Ro, R100 and Rt represent the resistances at 0°C, 100°C and t°C
respectively.
Using Eqn. (i), Eqn. (ii) can be written as
at + bt2
_____________
tpt = × 100 (iii)
100a + (1002)b
Then the true temperature can be given by
at + bt2
t – tpt = t –
[
_____________
100a + (1002)b ]
× 100

100at + (1002)bt – 100at – 100bt2


= ____________________________
100a + (1002)b
(1002)bt – 100bt2
= ______________
100 (a + 100b)
(1002)b ____ t2
= ________
[
t
– ____2
a + 100b 100 100 ]
[ t
( )]
t
= d ____ – ____
100 100
2
(iv)

(1002)b
________
where d=
a + 100b

Substituting the values of a and b we get d = 1.4929. Substituting the values of


Ro, R100 and Rt in Eqn. (ii), we get
tpt = 300°C

Since the rhs of Eqn. (iv) contains t, the true temperature can be evaluated by
the method of iteration. Thus t = 309°C.

QUESTIONS
1. Explain the terms intensive and extensive properties of a system.
2. Explain the terms ‘closed’ and ‘open’ system. What do you understand
by isothermal process, isobaric process, isochoric process and adiabatic
process?
3. Describe Joule’s experiment for determination of mechanical equivalent
of heat.
Basic Ideas 1.37

4. Describe a constant – volume gas thermometer. How it can be used for the
measurement of temperature of a gas inside a bulb?
5. Explain how the ideal gas equation gets modified in case of a real gas.
6. Why does the behaviour of a real gas differ from an ideal gas? Deduce the
van der Waals equation of state for the real gases.
7. Discuss the nature of van der Waals forces.
8. The volume of 1 mol of ideal gas at 0°C is 1 litre. Calculate its pressure.
If the gas obeys van der Waals equation of state and van der Waals gas
constants are a = 0.37 Nm4/mole2, b = 43 cm3/mole, find the pressure.
[Ans. 2.27 × 106 N/m2, 2 × 106 N/m2]
9. Write short notes on
(a) Avogadro’s number (b) Universal gas constant
(c) Platinum resistance thermometer (d) van der Waals interaction
(e) Triple point cell

OBJECTIVE TYPE QUESTIONS


1. The relationship between the heat energy and mechanical energy is studied
by
(a) Electronics (b) Ideal gas
(c) Thermodynamics (d) None of these
2. A vessel contains N molecules of a gas. If the number of molecules of gas
are made 2N, its pressure becomes
(a) double (b) 4 times
(c) halved (d) 1/4th
3. Suppose the temperature t be the linear function of the thermodynamic
property X. If Xo, X100 and Xt be the values of X at 0°C, 100°C and t°C
respectively than t°C is
Xt – X Xt – Xo
(a) t°C = _______ (b) t°C = ________ × 100
X100 – Xo X100 – Xo
Xo – Xt Xo – X
(c) t°C = _______ × 100 (d) t°C = ________ × 50
X100 – X X100 – Xo
4. The temperature of absolute zero is
(a) 273°C (b) 276°C
(c) – 273.16°C (d) – 274.17°C
5. When the solid, liquid and vapour coexist in equilibrium, then the system
is at the
1.38 Heat and Thermodynamics

(a) single point (b) double point


(c) triple point (d) none of these
6. The suitable metal for the resistance thermometer is
(a) pure platinum (b) copper
(c) iron (d) silver
7. A gas which obeys the Boyle’s and Charles laws is known as
(a) perfect gas (b) gas
(c) temperature (d) none of these
8. In the perfect gas equation PV = RT, V is the volume of
(a) 1 g gas (b) 1 litre gas
(c) 1 mole of gas (d) whole amount of gas
9. At a constant volume, the pressure of a given mass of a gas increases or
decreases by
(a) 1/273 of its pressure at 0°C (b) – 1/273 of its temperature
(c) 273 of its pressure at 0°C (d) both (a) and (b)
10. How many atoms or molecules are there in a mole
(a) NA = 6.02 × 1023 mole–1 (b) NA = 6.02 × 1022 mole–1
(c) NA = 6.02 × 1021 mole–1 (d) NA = 6.23 × 1023 mole–1
11. Ideal gas law is
(a) pV = T (b) p = VT
(c) pV = nRT (d) T = np

Answers
1. (c); 2. (a); 3. (b); 4. (c); 5. (c); 6. (a); 7. (a);
8. (c); 9. (a); 10. (a); 11. (c)
2
Chapter

Kinetic Theory

The matter is made up of molecules, which in turn, are composed of atoms. It


may be a matter of surprise that universal acceptance of the existence of atoms
by the scientific community did not occur until the early 1900’s. But today the
hypothesis that atoms exist is so essential to our understanding of the nature of
the world around us that the Nobel laureate physicist Richard Feynman could
state “If all scientific knowledge were to be destroyed, I would hope that the
knowledge that atoms exist might be spared”.

2.1 BROWNIAN MOTION


In solids, when heated, the molecules/atoms vibrate about their equilibrium
portions. In liquids and gases also, the molecules are undergoing ceaseless random
motion. The direct evidence of molecular motion was first found by British
Botanist Robert Brown in 1828. While studying very tiny spores suspended in
water, through a high power microscope, he observed that the spores were in
continuous chaotic motion. The motion is readily observed in colloidal solutions
under high power microscope. The suspended particles are seen to move entirely
haphazardly. This ceaseless irregular motion is known as the Brownian motion.
Experiments have shown that the nature of Brownian movements of suspended
particles depends on the properties of liquid or gas in which the particles are
suspended. The motion becomes vigorous when the temperature is increased and
a less viscous medium is chosen.
The only quantitative investigations of Brownian motion of particles were
made by Perrin in 1909. He found that the (suspended) particles do not settle
entirely to the bottom of the vessel but remain distributed. He verified that the
mean distribution of the particles was exponential under the gravitational field
as proportional to e– mgh/kT. Prior to his investigations, the following facts were
established:
1. The motions are completely irregular and random. The motion are
independent of the location of the particles and no two particles move in
the same direction.
2. They are independent of shaking of the vessel.
2.2 Heat and Thermodynamics

3. The lower the viscosity, the faster is the motion.


4. The smaller the particles, the greater the motion.
5. The motion are continuous and perpetual.
6. The molecules of same sizes move with equal speed at the same
temperature.

2.2 THEORIES OF BROWNIAN MOTION


The quantitative explanation of the Brownian motion was given by A. Einstein
and independently by M. Smoluchowski in 1906. We now give the derivation of
the Einstein formula as it was worked out by Langevin in 1908.

2.2.1 Langevin’s Theory of Brownian Motion


Langevin proceeded from the assumption that the force on a suspended particle of
mass m could be considered of two kinds:
(i) frictional force proportional to velocity
(ii) all other external influences of the surrounding fluid.
For the motion of a particle in any specified but arbitrary direction which we
take along the x-axis, the Langevin equation will be of the form:
m =–f +X (1)
where – f is the frictional force and X is the combined effect of all other influences.
Multiplying Eqn. (1) by x, we get
m = – f x + Xx (2)
1 d
Now, x = __ __ (x2)
2 dt
2
and
d
__
dt [ ]
1 d d(x )
( x) = __ __ ____
2 dt dt
2
or [ ]
1 d d(x )
x = __ __ ____ –
2 dt dt
2

\ Equation (2) becomes


2
f d(x2)
m __
__
[ ]
____) – m
d d(x
2 dt dt
2
= – __ ____ + Xx
2 dt
(3)

We form such an equation for each particle which is in the fluid and take the
mean of these expressions for all particles:
________ ____
2 2
m __
__
[ ]
2 dt dt
___
____) – m 2 = – __f d(x
d d(x
2 dt
___
____) + Xx
Kinetic Theory 2.3

___ As X varies in a completely irregular manner, one assumes that the mean value
Xx vanishes.
Also from the equipartition theorem:
___
2
m = kT
____
2
d(x d ( __2 )
____) = __
So, using x
dt dt
we get
______ ____
2 2
m __
__
2 dt dt[ ]
d d(x
____) + __
1 d(x )
f ____- = kT
2 dt
____
(4)

2
d(x
____) = u
Let
dt
__ du __
m ___ 1
Then + fu = kT
2 dt 2
The general solution for u is
2 ft
u = kT __ + C exp – __
f m ( )
where C is an integration constant. Now, as m is very small, the ratio f/m is very
large, so the exponential term has no influence after the first extremely small time
interval: ____
d(x2) 2
u = ____ = kT __ (5)
dt f
Integration from t = 0 to t = t, we have
__ ___
2
x2 – x02 = kT __ t
f
If now, we set x0 = 0 at t = 0 and write Dx2 instead of x2, then
__ 2
Dx2 = kT __ t (6)
f
The physical significance of f is that it is a force per unit velocity which can
be derived from Stokes law for the viscous force acting on a sphere of radius r
moving with velocity v in a fluid of viscosity h:
force F = 6phrv
F
Therefore, f = __
v = 6phr (7)

Combining Eqns. (6) and (7), we have


___ __
___ ___ __ 2
__
÷
2
÷Dx = ÷kT ÷t
f
2.4 Heat and Thermodynamics

___
___ ___ __ 1
or ÷Dx2 = ÷kT ÷t × ______
_____ (8)
÷3phr
___
R t
or Dx2 = __ T _____ (9)
N 3phr
where k ∫ R/N.
Here t is the time taken for random walk and N is the number of steps of random
walk. Equation (9) is the Einstein formula for the mean square displacement.
Since Brownian motion is brought about by molecular collision, this is of great
importance for the verification of the kinetic theory.

2.2.2 Einstein’s Theory of Translational Brownian Motion


The physical nature of Brownian motion is actually given by Einstein’s theory. It
is evident that the Brownian particles on account of their random motions, tend to
diffuse into the medium in course of time and therefore the diffusion coefficient
must be related to the Brownian motion.
The diffusion coefficient was calculated by Einstein in two ways:
1. from the random motion of the suspended particles.
2. from the osmotic pressure difference in different parts caused due to
difference in concentration of suspended particles. This concentration
difference leads to the phenomenon of diffusion.___
Let the diffusion coefficient be D. Let the value of ( Dx2 ) for the time t be D2.
Let us assume that the molecular concentration is not uniform but has a gradient
along the x-axis causing diffusion of the particles. For simplicity we suppose that
every particle suffers the displacement D in time t.
Now, imagine a cylinder full of molecules. The axis of the cylinder is parallel
to the x-axis and its end faces S1 and S2 are separated by D. Let the molecular
concentration at these ends be n1 and n2 respectively and let A denote the cross-
sectional area of the cylinder. The number of particles crossing the surface S1
in time t is 1/2 n1 DA, because half of the particles contained in the cylinder of
volume DA will cross S1 from the opposite side in time t. Similarly, the number
1
crossing S2 in opposite direction is __ n2 DA.
2
From the definition of diffusion coefficient
dn 1
– D ___ tA = __ (n1 – n2) DA (1)
dx 2
dn
where – ___ is the concentration gradient.
dx
dn
But – D ___ = n1 – n2
dx
Kinetic Theory 2.5

dn 1
or – ___ = __ (n1 – n2) (2)
dx D
therefore,
D
__ 1
(n – n2) t = __ (n1 – n2) D
D 1 2
D2
or D = ___ (3)
2t
Now, we calculate the coefficient of diffusion from the osmotic pressure
difference. Let the osmotic pressure acting on surface S1 and S2 be p1 and p2.
Then from gas laws
p1 = n1kT, p2 = n2kT (4)
Hence in the cylinder, there acts a force along x given by
(p1 – p2) A = (n1 – n2) kTA (5)
This force acts on nAD number of particles inside the cylinder, n being the
mean concentration. Thus the force acting on a single particle is
(n1 – n2) n1 – n2 kT
f = _______ kTA = ______
n
___
(nDA) D
dn
kT ___
= – ___
n dx (6)

(using Eqn. (2)).


and from Stoke’s law, the particle being suspended in viscous medium will move
with a steady velocity v
dn
kT ___
f = 6phrv = – ___
n dx (7)

– kT dn
or nv = _____ ___
6phr dx
This is the number of particles moving along the positive x-direction, per
second per unit area. Also, from the definition of diffusion coefficient

RT 1
D = ___ . _____ (8)
NA 6phR
Now, equating the two values of diffusion coefficient (from Eqns. (3) and (8)),
we get
D2 ___
___ 1
RT _____
= .
2t NA 6phr

RT 1
or D2 = ___ _____ t (9)
NA 3phr
2.6 Heat and Thermodynamics

Equations (3) and (8) have been verified experimentally. Brillouin measured the
coefficient of diffusion D and with the help of Eqn. (8) calculated the Avogadro’s
number NA. Equation (9) is the Einstein’s equation for the displacement D in
time t.

2.3 VERTICAL DISTRIBUTION OF BROWNIAN PARTICLES


AND DETERMINATION OF AVOGADRO’S NUMBER
The Brownian particles behave like gas molecules of large size. These particles
form a gas, which is in equilibrium under the action of the external field of earth’s
gravitation. The Brownian particles suspended in a liquid are in thermal motion
and experience a force of gravitational attraction of the earth. Therefore, the
distribution of these particles in the liquid will be similar to that of gas molecules
of the earth’s atmosphere. Let us now calculate the law of distribution in an
isothermal vertical column of the gas composed of Brownian particles.
Consider the vertical column of the gas. Let the pressure at a height Z be P.
When the height changes by dz, the pressure changes by dP. If r be the density of
the gas at height z, then force due to gravity acting on the layer is rgdz. Hence in
equilibrium
P = (P + dP) + rgdz
or dP = – rgdz (1)

where g = acceleration due to gravity.


Let n be the number of Brownian particles per unit volume and m the mass of
each particle,
\ r = mn
For a perfect gas
P = nkT
Consequently,
kTdn = – nmg dz (2a)
dn
___ mg
___
or n = – kT dz (2b)

Integrating, we get
mg
[
n = C exp – ___ z
kT ] (3)

where C is a constant of integration. If no be the number density of the Brownian


particles at a height zo, then the constant C can be determined from the condition
n = no at z = zo. Thus, we get
Kinetic Theory 2.7

mg
[
C = no exp ___ zo
kT ] (4)

Consequently,
mg
[
n = no exp – ___ (z – zo)
kT ]
[NA mg
= no exp – ______ (z – zo)
RT ] (5)

where NA = Avogadro’s number and R = NAk is the universal gas constant. The
above equation states that the number density of the Brownian particles decreases
with increase of height according to an exponential law.
From Eqn. (5), we have
no
RT
mg(z – zo) n( )
NA = _________ ln __ (6)

If r and r¢ be the densities of the suspended particles and the liquid, then the
effective mass m of the suspended particle is
4
m = __ p (s/2)3 (r – r¢)
3
1
= __ p s3 (r – r¢) (7)
6
where s is the diameter of the particle. Hence
no
3
6RT
N = _________________
p s (r – r¢) (z – zo)
ln __
n ( ) (8)

This is a convenient method for determining the Avogadro number, provided


the density r and diameter s of the suspended particles are known. Taking the
photographs of the suspended particles and counting the number of particles, the
value of n end no can be measured. Perrin calculated the Avogadro number as
NA = 6.82 × 1023 mol–1
which is slightly higher than the present value.

2.4 IDEAL GAS MACROSCOPIC DESCRIPTION


Of the three states of aggregation of matter, the gaseous state is the simplest one
due to the fact that the forces acting between the molecules are extremely small
and under certain conditions are negligible.
The gas in which the molecular forces are completely disregardable, is termed
an ideal gas. For simplicity, we also disregard the dimensions of the molecules,
particularly, while discussing the macroscopic properties of a gas i.e., pressure,
2.8 Heat and Thermodynamics

volume, temperature, internal energy and entropy. The picture of chaotic (random)
motion of gas molecules will be incomplete if we do not consider the collisions
of gas molecules with one another and with the walls of the container. Thus, due
to the aggregate effect of impact with the container walls, a pressure is exerted on
the walls.

2.5 KINETIC MODEL OF GASES


Basic postulates of the kinetic theory of gases are as follows:
(i) The molecules are in motion, and as they are material bodies, Newton’s
laws of motion may be applied.
(ii) A molecule moves in a straight line between two successive collisions
and the average distance traversed between the two collisions is called its
mean free path.
(iii) The gas is highly rarefied and the dimensions of molecules are infinitely
small so that the collisions among them are negligible.
(iv) Even if incessant collisions of the molecules are considered, in steady
state, the collisions do not affect the molecular density in any element of
volume of the gas.
(v) The molecules behave as perfectly elastic hard spheres and no appreciable
force of attraction or repulsion is supposed to be exerted between them.
(vi) The time during which a collision lasts is negligibly small compared with
the time required to cover the mean free path.

2.5.1 Pressure, Temperature and RMS Speed


Suppose n moles of an ideal gas is confined in a cubical box of edge L (volume V)
as in Fig. 2.1. The walls of the box are held at temperature T. We would now
derive the relation between the pressure P exerted by the gas on the walls and the
speeds of the molecules.
We, for the moment, ignore collisions
of molecules with one another and
consider only elastic collisions with the
walls.
Figure 2.1 shows a typical_ gas

molecule of mass m and velocity v that
is about to collide with the shaded wall.
Since the collision is assumed to be
elastic, when this molecule collides with Fig. 2.1 Cubical box of edge L containing n
the shaded wall, the only component moles of_an ideal gas. A molecule of mass m and

of its velocity that is changed is the velocity v is about to collide with the shaded wall.
A normal to the wall is shown
x-component, and that is reversed in
Kinetic Theory 2.9

direction i.e., the only change in the particles momentum is along the x-axis and
is given by
Dpx = (– mvx) – (mvx) = – 2mvx (1)
and the momentum Dpx delivered to the wall by the molecule during the collision
is + 2mvx.
The molecule of Fig. 2.1 will hit the shaded wall repeatedly. The time Dt
between collisions is the time the molecule takes to travel to the opposite wall
and back again (i.e., a distance of 2L) at speed vx. Therefore
2L
Dt = ___
v (2)
x
(Note that the above result holds even if the molecule happens to bounce off any
of the other walls along the way, as those walls are parallel to x and so cannot
change vx). Consequently, the average rate at which momentum is delivered to the
shaded wall by this single molecule is
Dpx 2mvx mvx2
___ = _____ = ____ (3)
Dt 2L/vx L
_›
_› dp
From Newton’s second law F = ___
dt
To find the total force on the wall, we need to add up the contributions of all the
molecules that strike the wall, allowing for the possibility of all having different
speeds (vx1, vx2, …, vxN). Since pressure is force per unit area, we have
pressure
2 2 2
Fx mvx1 /L + mvx2 /L + … + mvxN /L
P= 2=__ __________________________
L 2
L
m
( )
= __3 (v x1
L
2
+ v 2x2 + … + v xN
2
) (4)

where N = total number of molecules in the box. Since N = (nNA), there are
nNA terms in the second parentheses on the r.h.s. of the above equation so, we
can replace this quantity by nNA(vx2)avg, where (vx2 )avg is the average value of the
square of the x-components of all the molecular speeds. Then Eqn. (4) becomes
mnNA 2
P = _____ (vx )avg (5)
L3

But mNA ∫ the molar mass M of the gas (i.e., mass of 1 mole of the gas) and
3
L = V (volume of the cube), therefore
Mn (vx2 )avg
P= _________ (6)
V
2.10 Heat and Thermodynamics

For any molecule


vx2 + vy2 + vz2 = v2 (7)
Because there are many molecules and all are moving in random directions,
1
therefore the average values of their velocity components are equal i.e., vx2 = __ v2.
3
Consequently
nM (v2)avg
P = ________ (8)
3V
The square root of (v2)avg is a kind of average speed called the root mean
square speed of the molecules and denoted as vrms i.e.,
______
2
÷(v )avg = vrms (9)
and
nMv2rms
P= _______ (where M = mNA) (10)
3V
This equation, based on the kinetic theory, tells us how the pressure of the gas
(a purely macroscopic quantity) depends on the speed of the molecules (a purely
microscopic quantity).
From the above equation, and using the ideal gas law (PV = nRT), we get
____
3RT
vrms = ____
÷
M
(11)

where T is the temperature and R, the universal gas constant.

2.5.2 Translational Kinetic Energy of a Gas Molecule


in Terms of Temperature
Consider a single molecule of an ideal gas moving around in a box of edge L.
We assume that its speed changes when it collides with other molecules. Its
translational kinetic energy is 1/2 mv2 at any instant. Average translational kinetic
energy over the time we watch it is

1 1
Kavg = __ m (v2)avg = __ mv2rms (12)
2 2
assuming that the average speed of the molecule during our observation is the
same as the average speed of all the molecules at any given time (provided, the
total energy of the gas remains constant). But
____
3RT
vrms = ____
÷
M
(13)
Kinetic Theory 2.11

Therefore,
1
( )
3RT
Kavg = __ m ____
2 M
(14)

Further,
M
__
(molar mass
________________
m = mass of a molecule = NA) (15)

the Avogadro number


Thus
3RT
Kavg = ____ (16)
2NA
R
Using ___ = k, the Boltzmann constant,
NA
we get
3
Kavg = __ kT (17)
2
__
Now, let C2 __denote the mean square velocity of the molecules. Then average
energy Kavg ( ∫ E ) can be written
__ 1 __ 3
E = __ mC2 = __ kT (18)
2 2
i.e. average kinetic energy per mole of the molecules of a perfect gas depends
only on the temperature T and is independent of pressure P, volume V or species
of the molecules. Another useful form of Eqn. (18) is obtained by multiplying it
with NA:
1 __2 __
__ 3
MC = RT ( mNA = M and kNA = R)
2 2
i.e. average gives a physical meaning to R: R is equal to two thirds of the total
translational kinetic energy of the molecules in one gram molecule of a gas at
T = 1°K. In other words R defines the size of the degree.
__
The product of E and the total number of molecules N equals the total random
(translational)__kinetic energy of the molecules. We can identify it with the internal
energy U ∫ NE. Thus
__ 1 __
3
U = NE = __ NmC2 = __ NkT
2 2
3 3
= __ nRT = __ PV (20)
2 2
2.12 Heat and Thermodynamics

We can regard Eqn. (18) as a definition of gas temperature on the kinetic energy
basis
__ at the microscopic level. An increase
__ T of a gas is equivalent to an increase
in E of its constituent molecules or E μ T.

2.6 SOME DEDUCTIONS FROM KINETIC THEORY


2.6.1 Boyle’s Law
We have
1 __ 3
__ NmC 2 = __ PV
2 2
(M = mNA and nNA = N)
1 __ 3
__ nMC 2 = __ PV (1)
2 2
(where M is the molar mass)
\ For a given quantity of gas, n is constant
1 __ 1
and PV = __ nMC 2 ∫ __ MC2 (2)
3 3
The C remains constant if the temperature T is constant. Therefore
PV = constant
at constant temperature. (Boyle’s Law).

3.6.2 Charle’s Law


From Section 2.5.2 Eqn. (20)
3
__ PV = U (1)
2
3
__ 3
PV = __ nRT (2)
2 2
or PV = (nR)T (3)
Thus, if the pressure is kept constant, the volume increases linearly with the
kinetic energy of molecules. Since kinetic energy itself is proportional to the
absolute temperature (U μ T), therefore

VμT
at constant pressure (Charle’s law)

2.6.3 Avogadro’s Law


From Eqn. (1), we have, pressure
Kinetic Theory 2.13

1 __ __
P = __V mNC 2 ( representing C 2 ∫ V rms
2
) (4)
3
__
( C 2 is the mean square velocity ).
Consider now two gases a and b at the same temperature T and pressure P.
Then for equal volumes of the two gases
___ ___
1 1
PV = __ ma Na Ca2 = __ mb Nb Cb2 (5)
3 3
If the two are at the same temperature, the mean kinetic energy of molecules of
the two gases will also be the same i.e.,
___ ___
1
__ 1
m C 2 = __ m C 2 (6)
2 a a 2 b b
From Eqns. (5) and (6),
Na = Nb
hence at the same pressure and temperature equal volumes of all gases have the
same number of molecules. This is Avogadro’s Law.

2.6.4 Pressure Law


According to pressure law, at a constant volume, the pressure of given mass of a
gas is directly proportional to its absolute temperature.
From kinetic theory, pressure of an ideal gas
1 __
P = ___ mNC 2
3V

2 N 1 __
= __ __ × __ mC 2 (7)
3V 2
(where N is the total no. of molecules in volume V).
1 __ 2
__
But mC μ T
2
1 __ 2
__
or mC = aT (8)
2
(where a is a constant of proportionality)
2N
\ P = __ __ × aT
3V

( )
2 Na
= __ ___ T
3 V
(9)

i.e., for a given mass of the gas (N = constant) and if v is constant P μ T.


This is the pressure law.
2.14 Heat and Thermodynamics

2.6.5 Perfect Gas Equation


Pressure of an ideal gas
1 __
P = ___ mN C 2 (10)
3V
If volume of one mole of gas is V, then, N ∫ NA (where N is the total no.
molecules). Hence, pressure of one mole of gas will be
1 __
P = ___ mNA C 2 (11)
3V
2 1 __
or PV = __ × __ mNA C2 (12)
3 2
where m = mass of a molecule.
But mean kinetic energy of one molecule is
1 __ 2
__ mC μ T
2
1 __ 2
__
or mC = aT (13)
2
(where a is a constant of proportionality)
2
\ PV = __ × NA aT (14a)
3
or PV = RT (14b)
where R = 2/3 NA a = NAk is the universal gas constant (k = Boltzmann const.)

2.6.6 Graham’s Law of Diffusion


According to this law, at the same pressure, the rate of diffusion of a gas is
inversely proportional to the square root of its density.
From kinetic theory, pressure
1 __
P = ___ mNC 2
3V
__
or
3PV 3P
C 2 = ____ = ___
mN r ( mN
___
V
= r, the density )
___ ___
__ 3P
vrms = ÷C 2 = ___
or
÷r
(15)

i.e., at a constant pressure


1__
vrms μ ___
÷r
Since, the rate of diffusion of a gas is directly proportional to the root
__
mean
square velocity of its molecule, therefore, rate of diffusion of gas μ 1/÷r .
Kinetic Theory 2.15

2.6.7 Dalton’s Law of Partial Pressures


According to this law, at a constant temperature, the pressure of mixture of gases
(not interacting with each other) is equal to the sum of partial pressures of its
component gases.
Let there be n1 molecules of a gas in volume V (each of mass m1 and mean
___
square velocity C12). Then by the kinetic theory
Pressure of this gas ___
1 m1 N1 C12
P1 = __ ________
3 V
Similarly, pressure of gas 2 ___
1 m2 N2 C22
P2 = __ ________
3 V
Pressure of gas 3 ___
1 m3 N3 C32
P3 = __ ________ (16)
3 V
Now, if these three gases are mixed together and the volume of the mixture is
V, then since the temperature of each gas is the same, therefore, the mean kinetic
energy of molecules of the three gases will be equal i.e.,
___ ___
1
__ 1
m1 C12 = __ m2 C22
2 2
___ __
1
= __ m3 C32 = E (say) (17)
2
Therefore,
2 N1 __
P1 = __ ___ E;
3 V
2 N2 __
P2 = __ ___ E;
3 V
2 N3 __
P3 = __ ___ E (18)
3 V
and pressure of mixture
2 N __
P = __ __ E (19)
3V
where N = N1 + N2 + N3
__
2 E
\ P = __ (N1 + N2 + N3) __
3 V
2 1N __ 2 2N __ 2 N3 __
= __ ___ E + __ ___ E + __ ___ E
3 V 3 V 3 V
2.16 Heat and Thermodynamics

or P = P1 + P2 + P3 (20)
This is the Dalton’s law of partial pressures.

2.7 THE INTERNAL ENERGY AND DEGREES OF FREEDOM


We know that the average translational kinetic energy per molecule of an ideal
monatomic gas is K = 3/2 kT. For such a gas, this is the only total internal energy
as the molecules of an ideal monatomic gas have no potential energy, they cannot
vibrate or rotate, so the vibrational or rational energy are zero. Consequently, the
total internal energy of n moles of an ideal monatomic gas is
Eint = (nNA)K (as they are nNA number of molecules)
3
= nNA × __ kT
2
3
or Eint = __ nRT ( kNA = R) (1)
2
where R is the universal (molar) gas constant. Thus if we change the internal
energy of a gas by doing work on it or transferring heat to it, its temperature will
change and energy change
3
DEint = __ nR DT (2)
2
Now, let us consider not the monatomic (point) molecule but a molecule
consisting of two point particles separated by a small distance. This is how we can
represent the molecule of a diatomic gas such as O2, N2 or CO (carbon monoxide).
Such a molecule can acquire kinetic energy
y
by rotating about its centre of mass, and
then, we need to consider contributions to
the internal energy from rotational motion
and translational motion.
The rotational kinetic energy of the
diatomic molecule (Fig. 2.2) can be written
as
1 1
Krot = __ Ix wx2 + __ Iy w y2 (3) x
2 2
where I is the rotational inertion of the z
molecule for rotation about a particular axis.
Fig. 2.2 A diatomic molecule consisting
For point masses no kinetic energy is of two atoms (considered to be point par-
associated with rotational about the z-axis ticles) with its axis along the z-axis of the
and Iz = 0. The total kinetic energy of the coordinate system
diatomic molecule is
Kinetic Theory 2.17

1 1 1 1
K = __ mvx2 + __ mvz2 + __ Ix wx2 + __ Iy wy2 (4)
2 2 2 2

To find the total energy of the gas, we must find average energy of a single
molecule and then multiply by the number of molecules.
The five terms in Eqn. (4) represent independent ways in which a molecule can
absorb energy and are called degrees of freedom i.e.,
Degrees of freedom are the number of independent ways that molecules can
possess energy.
A monatomic molecule has three degrees of freedom. A diatomic molecule has
five degrees of freedom, (three translational and two rotational degrees).
Now, according to Maxwell’s theorem on equipartition of energy (see next
article):
When the number of molecules is large, the average energy per molecule
1
is __ kT for each independent degree of freedom.
2
3
Thus, for a monatomic ideal gas, the average energy per molecule is __ kT
2
( 1
)
3 degrees of freedom × __ kT and for NA molecules:
2

3
Eint = NA × __ kT
2
3
= __ nRT (monatomic gas)
2
For a diatomic gas,
5
Eint = NA × __ kT
2
5
= __ nRT (diatomic gas) (5)
2
A polyatomic gas (more than two atoms per molecule) generally has three
possible axes of rotation (unless the three atoms be in a straight line, as for CO2).
1
Then the kinetic energy of a single molecule would have a sixth term __ Iz wz2
2
and

( )
6
Eint = NA × __ kT
2

= 3nRT (polyatomic gas)


2.18 Heat and Thermodynamics

So far, we have considered only the contributions of translational or rotational


kinetic energy to the internal energy of the gas. Other kinds of energy may also
contribute. For example, a diatomic is also free to vibrate (imagine two point
atoms connected by a spring). Thus, there will be two additional contributions
due to potential energy of the spring and kinetic energy of oscillating atoms) then
diatomic molecule will have 7 (= 3 + 2 + 2) degrees of freedom. For polyatomic
molecules, the number of vibrational terms can be greater than 2. However,
vibrational modes are apparent only at high temperatures.

2.8 PRINCIPLE OF EQUIPARTITION OF ENERGY


Consider a system described classically in terms of f coordinates, q1, q2, …, qf and
f corresponding moment a p1, p2, …, pf. Its energy E is then a function of these
variables.
E = E (q1, …, qf, p1, …, pf) (1)
This energy if often of the form
E = Ei (pi) + E¢ (q1, …, qf, p1, …, pi – 1, pi + 1, …, pf) (2)
where Ei is the function of the particular momenta pi and E¢ depends on all
coordinates and momenta except pi (This functional form may arise because
kinetic energy of a particle depends only on momentum while its potential energy
depends only on its position).
Suppose the system under consideration is in thermal equilibrium with a heat
reservoir at an absolute temperature T. What then is the mean value of the energy
contribution Ei? Assuming classical statistics of identical but distinguishable
particles, the mean value of energy Ei is given by

__ Úe– bE(q1 … pf) Ei dq1 … dpf


Ei = ___________________ (3)
Úe– bE(q1 … pf) dq1 … dpf
( 1
)
Here, e– bE, b = ___ is known as the Boltzmann factor, which is due to classical
kT
particles obeying the Maxwell Boltzmann statistics.
In Eqn. (3), the integrals extend over all possible values of all the coordinates
q1 … qf and all the momenta p1 … pf. Using Eqn. (2), Eqn. (3) becomes

__ Úe– b(Ei + E¢) Ei dq1 … dpf


Ei = __________________
Úe– b( Ei + E¢) dq1 … dpf
__ e– bE Ei dpi Ú¢e– bE¢ dq1 … dpf
Ú_______________________
i

or Ei = (4)
Úe– bEi dpi Ú¢e– bE¢ dq1 … dpf
Kinetic Theory 2.19

The primes on the last integral signs in the numerator and denominator indicate
that these integrals extend over all the coordinate q and momenta p except pi.
Thus,

__ Úe– bEi Ei dpi


__________
Ei =
Úe– bEi dpi

(
– ___ Úe– bEi dpi
∂b
)
= ______________ (5)
– bE
Úe i dpi
+•

= – ___ ln
∂b [Ú –•
e– bEi dpi ] (6)

Here, the explicit limits in the integral reflect the fact that the momentum pi can
assume all possible values from – • to + •.
Consider now that Ei be a quadratic function of pi, as it would be if it represents
kinetic energy i.e.,

Ei = bpi2 (7)
where b is some constant. Then,
+• +•
2
– bEi
Ú e dpi = Ú e– bbpi dpi (8)
–• –•
__
Let y ∫ ÷b pi, then

+• +•
1__
( 1__
)
2
Ú – bEi
e dpi = ___ Ú e– by dy; dpi = ___ dy
–• ÷b – • ÷b
Hence,
+• +•
ln Ú
–•
– bEi
e
1
dpi = – __ ln b + ln
2 [Ú–•
2
e– by dy ] (9)

From Eqns. (6) and (9), we get


__ ∂
(1
Ei = – ___ – __ ln b
∂b 2 )
( the second term on RHS of Eqn. (9) does not involve b and so does not
contribute)
__ 1 1
or Ei = ___ = __ kT (10)
2b 2
2.20 Heat and Thermodynamics

If the functional forms of Eqns. (1) and (7) had been the same (except for
involving qi instead of pi) then all our previous arguments would be identical and
would thus lead again to Eqn. (10). Hence, we arrive at the following statement
known as the equipartition theorem:
“If a system described by classical statistical mechanics is in equibrium at
the absolute temperature T, every independent quadratic term in its energy has a
mean value equal to 1/2 kT.”
This is of great importance in thermodynamics for instance, due to this law, the
mixtures of ideal gases obeys the Dalton’s law of partial pressures.

2.9 MAXWELL‐BOLTZMANN LAW OF DISTRIBUTION OF


VELOCITIES IN AN IDEAL GAS
According to kinetic theory, each gas is made up of molecules. The molecules in a
gas move continuously in all the possible directions in a random manner and they
collide with themselves as well as with the walls of the container. In each collision,
the speed and direction of motion of the molecules changes i.e., velocities change
and it is quite probable to get molecules of any possible velocity at any instant.
Maxwell and Boltzmann arrived at a statistical law for the distribution of velocities
amongst the molecules of the gas. Following assumptions were made:
(i) Velocities can range between – • and + •.
(ii) In each volume element, the average number of molecules per unit volume
remains the same at any instant.
(iii) In each volume element, the velocity of molecules can have each possible
magnitude and direction whose distribution neither depends on the position
of volume element nor depends on the molecular collisions.
(iv) In the steady state, the number of molecules in different ranges of velocity
remains constant at a constant temperature although the velocity of each
independent molecule can vary.
Consider a perfect gas enclosed in a vessel of volume V at an absolute
temperature T. Let the number of molecules per unit volume be N. A molecule
whose speed at any instant is C and has the velocity components vx, vy, vz, then
vy
C 2 = vx2 + vy2 + vz2 (1)
P
We can now consider a velocity space having dvy dvz
vx, vy, vz as axes and velocity distribution functions dvx
c
f (vx), f (vy) and f (vz) such that f (vx) is independent
of vy and vz and f (vy) is independent of vx and vz; vx
and f (vz) is independent of vx and vy (Fig. 2.3). vz

Fig. 2.3
Kinetic Theory 2.21

We are to find the number of molecules whose velocity components are in


the range vx and vx + dvx, vy and vy + dvy and vz and vz + dvz. Obviously, such
molecules are represented by a parallelopiped in the vicinity of point p(vx, vy, vz)
and having sides dvx, dvy, dvz.
Let N f(vx) dvx = number of gas molecules in the velocity range vx and vx + dvx.
Similarly, we have N f(vy) dvy and N f(vz) dvz. Then, the number of molecules with
velocity components in the range vx and vx + dvx, vy and vy + dvy and vz and vz +
dvz will be

dN = N f(vx) N f(vy) N f(vz) dvx dvy dvz (2)


A molecule of a monatomic gas has only the translational motion for which the
number of degrees of freedom is 3, hence to describe its state we need to consider
a 6 dimensional phase space in which volume of each phase cell is h3 where h is
Planck’s constant. In this phase space, the number of phase cells in the volume
element dx dy dz dpx dpy dpz in between the position coordinates x, x + dx; y, y +
dy; z, z + dz and momentum coordinates px, px + dpx; py, py + dpy; pz, pz + dpz is

dx dy dz dpx dpy dpz


g = _________________ (3)
h3
and translational energy of the molecule is

px2 + py2 + pz2 p2


E = ___________ ∫ ___ (4)
2m 2m
Now, the probability P of finding the molecules in a given volume element of
phase space is
P μ ge– bE
1
where b = ___ (k = Boltzmann constant).
kT
dx dy dz dpx dpy dpz
P μ _________________
2
or 3
× e– p /(2mkT) (5)
h
Hence, the probability of finding the molecules in the momentum range px,
px + dpx; py, py + dpy; pz, pz + dpz inside the entire vessel will be
CV – (px2 + py2 + pz2)/(2mkT)
P(p)dp = ___ e dpx dpy dpz (6)
h3

( Ú Ú Ú dx dy dz = V )
C is a constant of proportionality.
2.22 Heat and Thermodynamics

The constant C is determined as follows, we have


+•

Ú P(p)dp = 1
–•

CV
Ú Ú Ú ___3 e– (p
2
+ py2 + pz2)/(2mkT)
or x
dpx dpy dpz = 1
h
+• +• +•
CV
___ 2 2 2
or Ú e– px /(2mkT) dpx Ú e– py /(2mkT) dpy Ú e– pz /(2mkT) dpz = 1
3 –•
h –• –•

From the standard integral


+• __
2 p
Ú e– ax dx = __ , ÷
–• a

we get
+• +• +•
– px2/(2mkT) – py2/(2mkT) 2
Ú e dpx = Ú e dpy = Ú e– pz /(2mkT) dpz
–• –• –•
________ ______
p
= ________ = ÷2pmkT
(1/2mkT)÷
h3
\ C = ___________ (7)
V(2pmkT)3/2
Substituting the value of constant C in Eqn. (6) we get

1
P(p)dp = ______
2pmkT ( ) 3/2 – (p 2 + p 2 + p 2)/(2mkT)
e x y z
dpx dpy dpz (8)

Substituting px = mvx, py = mvy, pz = mvz, we have

dpx dpy dpz = m3 dvx dvy dvz


Then from Eqn. (8) we have the probability of finding the molecules in the
velocity range (vx, vy, vz) to (vx + dvx, vy + dvy, vz + dvz) as

1
f(vx) f(vy) f(vz) dvx dvy dvz = ______
2pmkT ( ) 3/2 – m(v 2 + v 2 + v 2)/(2kT)
e x y z
× m3 dvx dvy dvz

(m2
= ______
2pmkT ) 3/2
e– m(vx
2
+ vy2 + vz2)
dvx dvy dvz (9)

m 3/2 – m(vx2 + vy2 + vz2)/(2kT)


\ dN = N _____
2pkT ( ) e dvx dvy dvz (10)

This is the Maxwell Boltzmann law of distribution of velocities.


Kinetic Theory 2.23

The number of molecules with x-component of velocity in the range 0 to vx is


m 1/2 – mvx2/(2kT)
(
N(vx) = N _____
2pkT ) e

Similarly

m 1/2 – mvy2/(2kT)
( )
N(vy) = N _____
2pkT
e

m 1/2 – mvz2/(2kT)
and N(v ) = N ( _____ )
z e
2pkT

2.9.1 Discussion on velocity Distribution Function


We can draw following conclusions from the above equation:
(i) Since N(– v ) = N(v ) N(vx)
x x
therefore a graph plotted for N(vx)vs. vx is (N(vx))max

symmetrical about the N(vx) axis (Fig. 2.4).


(ii) When vx = 0, N(vx) = maximum. Thus the
number of molecules with zero velocity
is maximum. In other words, the most
probably velocity (vx)m = 0 and
– ve O Vx +ve
m
| |
N(vx)
max (
= _____
2pkT ) 1/2
N Fig. 2.4 Plot between N(vx) and vx

As vx increases, the value of N(vx) decreases exponentially. At vx = •, N(v )


x
= 0.
(iii) The number of molecules is 1/e times the number of molecules at vx = 0
when

1
N(v ) = __
x e N(v ) | x |max

m 1/2 – mvx2/(2kT) m
1 _____
or (
N _____
2pkT ) e = __ (
e 2pkT ) 1/2
N
2
or e– mvx /(2kT) = e–1
______
or vx = ± ÷2kT/m

(iv) On increasing the temperature of gas, the number of molecules at the


maximum probably velocity vx = 0 decreases and the exponential fall
2.24 Heat and Thermodynamics

becomes slow, i.e., distribution N (Vx)


curve spreads (Fig. 2.5).
(v) The velocity distribution is different T
for different gases even at the same
temperature. On increasing the
mass of molecule (e.g., for a heavier
gas), the number of molecules 3T
corresponding to maximum
probable velocity vx = 0 increases
and the exponential fall becomes O VX
rapid i.e., the distribution curve
Fig. 2.5 Velocity distribution curve at two
contracts.
different temperatures
(vi) Average value of vx is

__
Area enclosed by N(v ) vs. vx curve with vx axis
vx or ·vxÒ = ______________________________________
x

Total number of molecules


+•

Ú v N
x (vx) dx
–__________

=
N
+•
m 1/2 – mvx2/(2kT)
vx N _____
Ú 2pkT ( e
=–__________________________

)
dvx
N
+•
m 2

(
= _____
2pkT ) 1/2
–•
Ú vx e– mvx /(2kT) dvx

But we have standard integral


+•
2
Ú xe– ax dx = 0
–•

\ ·vxÒ = 0
Similarly ·vyÒ = 0 and ·vzÒ = 0
(vii) Root Mean Square Value of vx
We know that
____
(vx)rms = ÷·vx2Ò
+•

Úvx2 N(v ) dvx


2 –____________
• x
But ·vx Ò =
N
Kinetic Theory 2.25

+•
m 1/2 – mvx2/(2kT)
or 2
·vx Ò =
Úvx2 N _____
2pkT
e(
–__________________________

dvx )
N
+•
m 2

(
= _____
2pkT ) 1/2
Ú
–•
e– mvx /(2kT) vx2 dvx

But, we have standard integral,


+•
2 1 ____
Ú x2 e– ax dx = ___ ÷p/a
–• 2a
+• _______
2 1 p
Ú vx2 e– mvx /(2kT) = _______ _______
( )÷
Hence m
–• 2 ____ m/(2kT)
2kT
_______
m 1 p
·vx Ò = _____
2
2pkT ( ) 1/2
× _______
2( )÷
m
____
_______
m/(2kT)
2kT
kT
= ___
m
_____
fi (vx)rms = ÷kT/m

Similarly,
_____
(vy)rms = ÷kT/m ;
_____
(vz)rms = ÷kT/m .

2.9.2 Maxwell-Boltzmann Law of Distribution of Speeds


Let c be the speed of the molecule which has vx, vy, vz as its velocity components.
Then
c2 = vx2 + vy2 + vz2 (1)
From the law of distribution of velocity

m 1/2 – m (vx2 + vy2 + vz2)/(2kT)


dN = N _____
2pkT ( ) e dvx dvy dvz

where dvx dvy dvz represents volume element in the velocity space. Since Eqn. (1)
is the equation of a sphere (of radius c) we may use spherical polar coordinates.
Then
dvx dvy dvz = c2 dc sin q dq df
2.26 Heat and Thermodynamics

and total volume of the shell between c and c + dc is


p 2p
2
= Ú c dc sin q dq Ú df
0 0

= 2pc2 dc (– cos q)p0 = 4pc2 dc

Consequently, the number of molecules with speeds in the range c and


c + dc is
m 3/2 – mc2/(2kT)
N(c) dc = N _____
2pkT ( ) e 4pc2 dc

m 3/2 2 – mc2/(2kT)
or N(c) dc = 4pN _____
2pkT ( ) c e dc

m 3/2 2 – mc2/(2kT)
or N(c) = 4pN ( _____ ) c e
2pkT
This is the Maxwell Boltzmann’s speed distribution function.

Most Probable Speed cm


It is the speed corresponding to which the number of molecules is maximum
(= N(c)) i.e.,

At c = cm, | N(c) | = maximum


d
___
or N |
dc (c) c = cm
=0

d m
or ___
dc { (
4pN _____
2pkT ) 3/2 2 – mc2/(2kT)
c e }c = cm
=0

m
) { 2ce –mc2/(2kT) 2mc
( )}
2
or (
4pN _____
2pkT
3/2
+ c2 e– mc /(2kT) – ____
2kT c = cm
=0

or ( mc2
1 – ____
2kT ) c = cm
=0

____
2kT
\ m ÷
cm = ____ (2)

Speed Distribution Function


When c = 0, N(c) = 0 i.e., in translational motion, no molecule has
2
the zero speed.
For lower values of c, the term c2 is more effective than e–mc /(2kT) therefore as
Kinetic Theory 2.27

c increases, the number of molecules


corresponding to it increases and for c = cm,
the number of molecules is maximum. On
2
further increasing c, the term e–mc /(2kT)
becomes more pronounced, and so, N(c) then
decreases exponentially with c. At c = •,
N(c) = 0 (Fig. 2.6).

Number of Molecules Corresponding


to Maximum Probable Speed (N(c))max
______
At c = cm = ÷2kT/m

( m
(N(c))max = 4pN _____
2pkT ) ( ____
3/2
m )e
2kT m2kT/(m2kT) Fig. 2.6 Maxwell Boltzmann speed
distribution function
_____
m
or (N(c))max = 4N _____ e– 1
÷
2pkT
(3)

As the temperature of gas increases, the maximum probable speed cm increases,


but the number of molecules corresponding to it decreases. The exponential fall
of N(c) with c becomes slow. But the area enclosed by the curve (with c-axis)
remains the same i.e., the distribution curve spreads (Fig. 2.7).

N(c)
4T

Fig. 2.7 Speed distribution curve at two different temperatures

Mean or Average Speed


Ú c N(c) dc
·cÒ = ________
0
N
2.28 Heat and Thermodynamics

Substituting the value of N(c), we get



1 m 3/2 2 – mc /(2kT) 2
·cÒ = __
N
Ú c × 4pN ( _____
2pkT )
c e dc
0

m 2

(
= 4p _____
2pkT )
3/2
Ú c3 e– mc /(2kT) dc
0

But from standard integral,



2 1
Ú x3 e– ax dx = ____2
0 2a
Hence,

2 1
Ú c3 e– mc /(2kT) dc = ________
m 2
0 2 ( 2kT )
____

m 1
·cÒ = 4p ( _____ ) ________
3/2
\ m
2pkT 2 ( ____ )
2

____ 2kT
8kT
or ·cÒ = ÷____
mp (4)
___
Root Mean Square Speed ÷·c2Ò

Ú c2 N(c) dc
·c2Ò = _________
0
N

1 m 3/2 2 – mc2/(2kT)
= __ Ú c2 × 4pN _____
N0 2pkT ( ) c e dc


m 2

2pkT(
= 4p _____ ) 3/2
Ú c4 e– mc /(2kT) dc
0

Now, from the standard integral


• __
2 3 p
Ú x4 e– ax dx = ____2 __ ÷
0 8a a
we get
• ______
2 3 p
Ú c4 e– mc /(2kT) dc = ________ ______
0
m 2
____
8 ( 2kT ) ÷( ) m
____
2kT
Kinetic Theory 2.29

3kT
= ____
m
____ ______
\ crms = ÷·c2Ò = ÷3kT/m (5)

Relation Between the Most Probable Speed, Average Speed


and R.M.S. Speed
Most probable speed
______
cm = ÷2kT/m
Average speed
_______
cav ∫ ·cÒ = ÷8kT/mp
R.M.S. speed
____ ______
crms ∫ ÷·c2Ò = ÷3kT/m
Hence,
cav
cm _____
___ crms ___
kT
__ = ___ = ____
__ = (6)
÷2 ÷8/p ÷3 m

2.10 MAXWELL BOLTZMANN’S ENERGY DISTRIBUTION LAW


Molecules in a gas are identical but distinguishable. They move randomly with all
possible speed in all possible directions. We are to find the number of molecules
in the energy range Œ and Œ + dŒ. In a monatomic gas, each molecule has only
translational motion.
Consider 1 mole of monatomic ideal gas in a box of volume V, at an absolute
temperature T. According to Maxwell-Boltzmann’s law, the number of molecules
in the energy range Πand Π+ dΠis

dN = Cg(dŒ) e– Œ/kT (7)

where g(dŒ) is the number of microstates in the energy range Œ and Œ + dŒ.
Since each molecule has only the translational motion (with 3 degrees of
freedom) therefore, we consider a six-dimensional phase space to express its
state. The volume of each phase cell is h3.

1
Now g(dŒ) = __3 Ú Ú Ú Ú Ú Ú dx dy dz dpx dpy dpz
h

where Ú Ú Ú dx dy dz = V (volume of position space)

and Ú Ú Ú dpx dpy dpz = volume of momentum space = V.


2.30 Heat and Thermodynamics

If momentum components of a molecule are px, py and pz and p is the total


momentum (of a molecule) then

p2 = px2 + py2 + pz2

This is the equation of a sphere of radius p in momentum space. Thus

Ú Ú Ú dpx dpy dpz = 4pp2 dp


V
and g(dŒ) = __3 × 4pp2 dp
h
_____
But p2 = 2mŒ or p = ÷2mŒ

___ ___
1 m
dp = ÷2m × __ × Œ– 1/2 dŒ = ___ dŒ
and
2 2Œ ÷
___
V m
g(dŒ) = __3 × 4p (2mŒ) × ___ dŒ

h 2Œ ÷
__
4÷2
= ____ pVm3/2 Œ1/2 dŒ
h3
Substituting this value in Eqn. (7), we get
__
C × 4÷2
dN = _______
3
pVm3/2 Œ1/2 × e– Œ/kT dŒ (8)
h
To find the value of constant C, we integrate the above equation for Πfrom 0
to •, therefore
__ •
4÷2 pVm3/2
_________
Ú dN ∫ N = C × Ú Œ1/2 e– Œ/kT dŒ (9)
h3 0
p2
But Π= ___, therefore
2m
pdp
dΠ= ____
m
__ •
4÷2 pVm3/2 ______
1 2
\ N = C × __________
3
× __ 3/2 Ú p2 e– p /(2mkT) dp (10)
h ÷2 m 0
• __
2 p
1 __
But Ú x2 e– ax dx = ___
4a a ÷
0
• _______
2 – p2/(2mkT) 1 p
fi Úp e dp = _________ _______
0
( 1
4 _____
2mkT
1
) ÷(
_____
2mkT )
Kinetic Theory 2.31

Nh3
fi C = ___________ (11)
V(2pmkT)3/2
Substituting the value of C in Eqn. (8)
__
Nh3 4 ÷2 Vm3/2 1/2 – Œ/kT
dN = ___________ × _____________ e e dŒ (12)
V(2pmkT)3/2 h3

or dN = 2pN ____
pkT ( )
1 3/2 1/2 – Œ/kT
Πe dΠ(13)

\ The probability for a molecule to be found in the energy range Œ and Œ + dŒ


is
dN 1
P(Œ) dŒ = ___ = 2p ____
N pkT ( ) 3/2
Œ1/2 e– Œ/kT dŒ (14)

and the probability for a molecule to have energy in range 0 to Πis


1
P(Œ) = 2p ____
pkT ( ) 3/2
Œ1/2 e– Œ/kT (15)

Most Probable Energy (Œm)


At Œ = Œm, P(Œ) is maximum

or [ ___
d

P ]
(Œ) Œ = Œ
m
=0

or
1
( )
2p ____
pkT
3/2 d [ 1/2 – Œ/kT ]
___

Œ e Œ = Œm = 0

or [ ____
1
__
2 Œ
÷
e – Œ/kT 1
– ___ Œ1/2 e– Œ/kT
kT ] Œ = Œm
=0

kT
fi Œm = ___ (16)
2
This is independent of the mass of the molecule.

Mean Energy of a Molecule


·ŒÒ = Ú Œ P(Œ) dŒ
0

( )
1
= Ú Œ × 2p ____
0 pkT
3/2
Œ1/2 e– Œ/kT dŒ

= 2p ( ____ ) Ú Œ
1 3/2 3/2 – Œ/kT
e dŒ
pkT 0
2.32 Heat and Thermodynamics

Let Œ1/2 = x or Œ = x2, therefore dŒ = 2x dx


Then
1
( )
·ŒÒ = 2p ____
pkT
3/2 2
Ú x3 e– x /kT (2xdx)
0

= 4p ( ____ )
1 3/2 2

pkT
Ú x4 e– x /kT dx
0

But from standard integral


• __
2 3 __p
Ú x4 e– ax dx = ____
2 a ÷
0 8a

2 3 ____
\ Ú x4 e– ax dx = _______2 × ÷pkT
0 8(1/kT)
____
Hence ( )
1 3/2 _______
·ŒÒ = 4p ____
pkT
×
3
8(1/kT)2
× ÷pkT

3
or ·ŒÒ = __ kT (17)
2
For 1 mole of gas
3
·ŒÒ = NA × __ kT
2
3
or ·ŒÒ = __ RT (18)
2
where NA is Avogadro number and R is universal gas constant.
Example 2.1 The molar mass M of oxygen is 0.0320 kg/mol.
(a) What is the average speed of oxygen gas molecules at T = 300 K?
(b) What is the root mean square speed at 300 K?
(c) What is the most probable speed at 300 K?
Solution
____
8RT
cavg = ____
(a)
pM ÷
(
8 × 8.31 J/mol.K × 300 K
= _____________________
p (0.0320 kg/mol) ) 1/2

= 445 m/s.
____
3RT
crms = ____
(b)
M ÷
_____________________
3 × 8.31 J/mol.K × 300 K
= _____________________ = 483 m/s.
÷ 0.0320 kg/mol
Kinetic Theory 2.33

____
2RT
cm = ____
(c)
÷M
_____________________
2 × 8.31 J/mol.K × 300 K
= _____________________
÷ 0.0320 kg/mol
= 395 m/s.

2.11 EXPERIMENTAL VERIFICATION OF


MAXWELL’S VELOCITY DISTRIBUTION LAW
In view of the fundamental importance of the Maxwell’s distribution law in kinetic
theory of gases, it was subjected many times to experimental verification. Many
attempts were made and some of these are considered as follows:

1. Stern’s Experiment
Maxwell distribution law of velocities has been verified by the experimental
arrangement due to Stern. The principle of Stern’s experiment can be explained
by Fig. 2.8, where L is a platinum wire coated with silver. The wire L serves as the
source of atoms, whose velocity is studied. When the wire is heated by an electric
current, it emits atomic silver in all directions. The wire L is surrounded by two
cylindrical diaphragms with narrow slits S1 and S2. The slits are parallel to the
wire. Through these slits, a stream of silver escapes and condenses on the plates P
and P¢. The whole apparatus is enclosed in a highly evacuated glass vessel, so that
the silver atoms may not suffer any collisions in space. The slits S1 and S2 and the
plates P and P¢ rotated together as a rigid body about the wire L as the axis.

Fig. 2.8 The principle of Stern’s experiment

When the entire system is at rest, the silver stream transverses along LO and
deposited at O. When the system is rotated at a high speed in a clock-wise as
shown in Fig. 2.8, the silver molecules will no longer strike the target at O but will
be displaced from O and deposited at some point A above O. The faster moving
molecules will condenses near O than slow ones. Thus the velocity spectrum
2.34 Heat and Thermodynamics

of silver molecules will be obtained. When the relative intensity of deposit is


measured with the help of a micro-photometer; the ratio of the number of
molecules with different velocities can be deducted and the Maxwell distribution
law is verified.
The result obtained by Stern are not very satisfactory. This is due to the
difficulty in retaining the perfect vacuum in the vessel because the spindle of the
rotting system 1 projected outside the vessel for coupling it to the deriving motor.
As a result the Maxwellian distribution law has been verified within about 15
percent.
This method has been improved by many scientists. One of the improvements
is due to Zartman and KO in 1930, which is described below.

2. Zartman and KO’s Experiment


Zartman and KO in 1930 have modified the Stern method to study the distribution
of velocities of molecules. The apparatus consists of an oven V with an opening
A (Fig. 2.9) S1 and S2 are two parallel slits. Above the slits, there is a cylindrical
drum D, which can be rotated in the vacuum about an axis passing through O. A
slit S3 is on one side of the drum and G is the glass plate mounted on the inside
surface of the drum opposite the slit S3.
Bismuth is taken as the experimental substance, which is heated and vaporized
in the oven. A molecular beam of Bismuth escaping through A is collimated by
the slits S1 and S2. When the drum
is stationary, the beam of molecule
entering into it through the slits S3,
Vacuum
strikes the glass plate at the same Pump
point. When the drum is rotated at a
high speed the molecules with very
G
high speed reach the glass plate G 0 D
first i.e., on the right end of G and
molecules with slower speed reach
the plate G the other end of G. After
a short time, sufficient quantity of
S3
Bismuth molecules is deposited on the S2
plate, whose density varies across G
according to the velocity distribution S1
of molecules. The thickness of the A
deposit i.e., the density distribution
is measured by a microphotometer. V

The results obtained by Zartman


are shown in Fig. 2.10, when the Fig. 2.9 Apparatus of Zartman and
relative intensity of deposit is plotted KO’s experiment
Kinetic Theory 2.35

against the displacement S (in cm). The circles denote the observed density of
deposit while the line represents the theoretical distribution on the basis of the
Maxwellian distribution law. The agreement is good which confirms the Maxwell
distribution law.
I/I0

s
Fig. 2.10 The relative intensity I/Io of deposit as a function of displacement S. The circles
represent the experimental values

3. Stern’s Improved Experiment


In 1947, Stern, Estermann and Simpson arranged a more precise experiment for
verifying the law of distribution of velocities. This experiment is based on the
method of molecular beam, where the free fall of the molecular is observed in the
gravitational field.
The experimental arrangement is schematically shown in Fig. 2.11. The
apparatus consists of an over V with a narrow horizontal opening A. Cesium is

v S
T
A

T≤

Fig. 2.11 Apparatus of Stern’s improved experiment


2.36 Heat and Thermodynamics

taken as the source of atoms, which is heated in the oven. A slit S is placed at a
distance of 1 metre from A, T is a thin tungsten wire placed at a distance of 1
metre from the slit s. It serves as a target A. The entire arrangement A, S and T are
along one strictly horizontal line. The entire arrangement is enclosed in a highly
evacuated chamber.
The Cesium atoms flow out of the oven through A. In the absence of gravitational
field, they would strike the target at T. However, due to gravitational field the atoms
travel along a parabola. The atoms emerging from A with a velocity horizontally
along the X-axis will not pass through the slit S and will not reach the target.
The atoms emerging from A at a small angle q, (as shown in Fig. 2.11) will pass
through S and strike the target.
The tungsten wire-target is heated by an electric current passing through it.
When Cesium atoms strike the wire-target, they get ionized. These positively
charged ions, leaving the target, get into the negatively charged cylinder
surrounding the target. Thus, an electric current of ions passes between the wire
and the cylinder, which can be measured with accuracy. The ionic current gives
the number of atoms hitting the target. Moving the target in a vertical direction
such as positions T ¢ and T ≤, the ionic current and hence the number of atoms
hitting the target is measured at different heights. We find the number of atoms
having different velocities-the atoms hitting at T¢ have velocity higher than those
at T ≤. This gives us the distribution of atoms with velocities. This is in complete
agreement with the Maxwell distribution law of velocities.

2.12 MEAN FREE PATH


If molecules were geometrical points, they
would not collide at all. Molecules, however, 10
are not points. They are more like hard elastic 12 5
spheres. Therefore, collisions occur.
If we could observe a single molecule 11 6
travelling through a gas, its path would look
something like that shown in Fig. 2.12. 2 8
Between collisions the molecule moves freely 9
with a constant speed along a straight line. 3
The distance between two successive 7
collisions, for example l2, 3, is called a free
path. The average length of these free paths is
4
called the mean free path l. If l1, 2, l2, 3, …,
etc. are the values of successive free paths, 1
_
l1, 2 + l2, 3 + … + lv – 1, v + lv, v + 1 = ct (1) Fig. 2.12 Path of a molecule
Kinetic Theory 2.37

_
where c is the mean speed of the molecule and t the total time in covering the free
paths, then
l1, 2 + l2, 3 + … + lv, v + 1 _ct
l ∫ _____________________
v = __
v (2)

Here v is the number of collisions suffered in time t.


Let d be the diameter of a hard sphere molecule. Any other molecule B, whose
centre is within a distance d from the molecule A under consideration, will collide
with A, Fig. 2.13. In other words, a collision occurs whenever the centre of another
molecule comes within the ‘sphere of influence’ of radius d drawn around the
centre of A. For our simplest model we assume that only A moves and all the other
gas molecules are instantaneously at rest.

Fig. 2.13 Number of molecular collisions


_
If A is moving with the average speed c, its ‘sphere of influence’ sweeps out
_
in time t a cylinder of base pd 2 and length ct. The quantity pd 2 is called the
collision cross section of the hard sphere model. If nv is the number of molecules
_
per unit volume, this cylinder will contain (pd 2 ct) nv molecules. Therefore, in all
2_
A experiences v = pd ctnv collisions and its mean free path is
_ _
ct _______
ct 1
l = __ _____
v = pd 2 _ctn = pd 2n (3)
1 v

In deriving Eqn. (3), we have assumed that A collides with other stationary
molecules. Actually, all the molecules are moving. Consequently, we should write
_ _
v = pd 2 crel tnv, where crel is the mean relative speed of A with respect to other
molecules.
A detailed calculation, taking___ into account the actual speed distribution of__
_
the molecules, yields crel = ÷2 c. In a simple picture, the origin of the factor ÷2
can be seen by considering the relative velocities of two molecules at the point
of collision, Fig. 2.14. The extreme cases are the head-on collision (a) and the
grazing collision
__ _ (b). The average case appears to be the 90° collision (c), for
_
which crel = ÷2 c. Therefore, Eqn. (3) becomes
2.38 Heat and Thermodynamics
_ _
ct ________
c
l = __
v = pd 2 _c n
rel v

1 0.707
= ________
__ = ______ (4)
÷2 p d 2 nv p d 2 nv

c c

c
2 c

c
c
crel = 2c crel = 0 crel = 2 c

(a) (b) (c)

Fig. 2.14 Relative speeds: (a) head-on collision, (b) grazing collision, and (c) right-angle collision

We have p = nv kT. Hence, l can be expressed as

kT
l = _______
__ (5)
÷2 p d 2p

Using nv = 3 × 1019 molecules/cm3 at NTP and d = 2 × 10–8 cm, we find


l = 2 × 10–5 cm which is slightly smaller than the wavelength of visible light.
Thus, at sea level (p = 760 mm-Hg) l ~ 10–5 cm for air molecules. At 100 km
above the earth (10–3 mm-Hg ª 10–6 atm) l = 100 cm.

2.13 TRANSPORT PHENOMENA


In problems like viscosity, heat conduction, or diffusion one is interested in the
transport of something (like momentum, thermal energy, or mass) from one part
of a gas to another. For this reason they are called transport phenomena.
The transport of momentum, for example, can be understood with the help of
the following analogy. Consider two trains 1 and 2 moving along parallel tracks in
same direction but at different speeds, Fig. 2.15. As they pass one another suppose
some passengers start jumping from 1 to 2 and from 2 to 1. Suppose v1 < v2 and
all passengers have equal mass m. Then the jump 2 Æ 1 is from the more rapidly
moving train to the slower one and therefore transports momentum of amount + m
(v2 – v1) in one jump. This tends to speed up the slower train 1 when the passenger
lands upon it from 2. The jump 1 Æ 2 from the slower train to the faster train, on
the other hand, tends to slow down 2.
Kinetic Theory 2.39

Fig. 2.15 Exchange of momentum

If n pairs of passengers are exchanged per unit time between the two trains,
there will be an average force + m (v2 – v1) n tending to speed up train 1 and an
average force – m (v2 – v1) n tending to slow down train 2. A distant observer who
is unable to see the exchange of passengers will feel that this result is due to some
frictional drag between the trains.
The mechanism by which one layer of flowing gas exerts a viscous drag on
an adjacent layer is exactly similar, the gas molecules taking the role of the
exchanged passengers.

2.14 GAS VISCOSITY


Consider a gas at constant temperature, moving in the x-direction with a velocity
which increases with z. We know that (i) a molecule changes in velocity only
on collision, and (ii) it moves on an average a distance l before it undergoes
collision. Therefore, the motion of the gas can be regarded to occur in parallel
layers such that adjacent layers with different velocities are separated by a
distance l, Fig. 2.16.

Fig. 2.16 Motion of a gas in layers


2.40 Heat and Thermodynamics

Let B be an arbitrary layer with a uniform velocity v in the x-direction and a


velocity gradient dv/dz. Then the layer C above it has a flow velocity v + l(dv/dz)
and the layer A below it has a flow velocity v – l(dv/dz).
For simplicity, we assume:
_
(a) that the flow velocity v is very small compared with the mean velocity c of
the gas molecules due to their random motion, and
(b) that all molecules which reach B have travelled exactly a distance l with
_
the mean speed c since their last collision took place.
A molecule of mass m reaching layer B from layer C brings horizontal
momentum

( dv
px(C) = m v + l ___
dz ) (1)

A molecule reaching B from A brings horizontal momentum

( dv
px(A) = m v – l ___
dz ) (2)

For molecules, moving at random, all directions are equally possible. On


average, one-third of all the molecules will be moving parallel to each of the
three coordinate axes. Moreover, on average, the number of molecules moving in
any one direction of given sign will be one-sixth of the total number.
Draw a cylinder RQ of height l and unit area of cross-section between the
layers C and B, Fig. 2.15. Its volume is l and contains nvl molecules, where nv is
1
the number of molecules per unit volume. Of these __ nvl are moving downward
6
_ _
with a speed c. Molecules will take time l c to travel from R to Q. In this time
all the molecules moving downward in the cylinder RQ will cross the unit area
Q. That is, the number of molecules crossing Q in downward direction in one
second is
1
__ nl
6 v
_____ 1 _
__
_ -= nvc (3)
l c 6

The same number will cross Q in the upward direction as can be evaluated by
drawing the cylinder PQ between the layers A and B.
1 _
The layer B itself will also discharge __ nvc molecules per second per unit
6
area towards both A and C layers. The resultant behaviour is summarized in
Table 2.1.
Kinetic Theory 2.41

Table 2.1 Molecules crossing unit area Q in the layer B per second.
Number Horizontal momentum

Entering from layer C


1 _
__
6 v
nc ( )
dv 1 _
m v + l ___ __ nvc
dz 6

Entering from layer A


1 _
__
6 v
nc ( )
dv 1 _
m v – l ___ __ nvc
dz 6
Leaving from layer B 1 _
__ 1 _
nc mv __ nvc
upwards 6 v 6
Leaving from layer B 1 _
__ 1 _
nc mv __ nvc
downwards 6 v 6
Net addition to layer B 0 0

We find that there is no accumulation of molecules or momentum in layer B.


However, there is a transport of momentum per unit area per second through the
layer B in the downward direction given by
1 _ __
__ 1 _ dv
p(C) (A)
x – px × nvc = mnvc l ___ ∫ F (4)
6 3 dz
This horizontal momentum per second is a horizontal force which is transmitted
through the gas to the top moving surface. In other words, this horizontal
momentum per unit area per second is equivalent to the shearing (tangential)
stress F.
The defining equation of viscosity is
dv
F = h ___
dz
where h is the coefficient of viscosity. By comparison, we have

1 _ 1 _
h = __ mnv c l = __ r c l (5)
3 3
__
Since l = ( ÷2 pnv d 2 )–1, we can write
_
mc
h = ______
__ (6)
÷2 p d 2
_
This relations can be used to compute molecular diameters, since m, c and h
can all be measured (or calculated).
_
From Eqn. (6), we conclude that the viscosity of a gas is proportional to c, that
1/2
is to T , and is independent of nv, that is, of the pressure. Both these conclusions
are in good agreement with experiments. Deviations occur at very low pressures
2.42 Heat and Thermodynamics

when l becomes comparable to the dimension of the container (l cannot increase


further by reduction of nv or pressure).

2.15 THERMAL CONDUCTIVITY


z
In this case the temperature T of a gas increases in
the z-direction. The situation can be experimentally T + l dT
C dZ

realized by confirming a gas between parallel walls, l


one of which is at a higher temperature. T
B
The temperature gradient is dT/dz. We again have l
layers separated by distance l. We can draw a figure dT
T – l dZ
A
similar to Fig. 2.16 or its simplified version shown
in Fig. 2.17. The heat flow per second per unit area x
across the layer B is
Fig. 2.17 Model for transport
dT
Q = K ___ of thermal energy
dz
where K is the thermal conductivity coefficient.
We assume that in a gas heat conduction is due to the transport of thermal
energy. The transport of mean thermal energy per unit area per unit time into
plane B from planes C and A is

( dT 1 _
)
Q(C) = mcv T + l ___ __ nvc from the upper layer C
dz 6

= mc ( T – l ___ ) __ n c from the lower layer A


dT 1 _
Q(A) v v
dz 6
where cv is the specific heat at constant volume per unit mass. The net transport of
thermal energy through the layer B per unit area per unit time is
1 _ dT
Q = Q(C) – Q(A) = __ nnc lmcv ___
3 dz
By comparison we see that the thermal conductivity is given by
1 _ 1
K = __ mnvc l cv = __ rc lcv (7)
3 3
Since l μ 1/r, K, like h, is independent of pressure and increases with
temperature.
From Eqns. (5) and (7), one obtains
K
__ = cv (8)
h
Experimentally also it is found that K/hcv is a constant, although its value is
different from unity. The observed ratio is between 1.4 and 2.5 for a very wide
range of gases, ever though the individual quantities K, h, cv vary much more.
Kinetic Theory 2.43

2.16 GAS SELF DIFFUSION


Suppose we have a concentration gradient in a z
gas. The molecules will move from the more n + l (dn1/dz)
concentrated to the less concentrated regions so C
l
that the property being transported is simply the B
n
gas molecules themselves. l
n – l (dn1/dz)
Let nv be the number of molecules per unit A
volume and dnv/dz the concentration gradient at x
the layer B, Fig. 2.18. The net number of molecules
Fig. 2.18 Concentration layers
transported across B per unit area and per unit time
is

n(C) (A)
v – nv =
1
__
6 ( dnv _ 1
) ( dnv _
nv + l ___ c – __ nv – l ___ c
dz 6 dz )
1 _ dnv
= __ lc ___ ∫ G
3 dz
The basic law of diffusion, due to Fick, is
dnv
G = D ___
dz
where D is the coefficient of self diffusion. Therefore
1 _ 1
3
1
D = __ lc = __ _____
3 pd nv
_
2( )_
c (9)

For air at NTP; d = 2.5 Å, l = 1000 Å, c = 45000 cm sec–1. This gives D to be


of the order of 1 molecule per cm2 per unit concentration gradient.
From Eqns. (5) and (9), we have
h
__ =r (10)
D
Thus, the single kinetic theory establishes general correlations between diverse
phenomena such as viscosity, thermal conductivity and self diffusion.
The results of mean free path treatments of the transport processes studied are
summarized in Table 2.2.

Table 2.2 Comparison of transport processes.


Process Transfer of Result cgs units of coefficient
_ 1 _
Viscosity Momentum, mc h = __ rcl g cm–1 sec–1
3
1 _ 1 _
Thermal conductivity Kinetic energy, __ mc2 K = __ rclcv crg cm–1 sec–1 degree–1
2 3
1 _
Diffusion Mass, m D = __ rc cm2 sec–1
3
2.44 Heat and Thermodynamics

Example 2.2 Assuming the Avogadro number NA = 6.02 × 1023 mol–1, universal
gas constant R = 8.31 Jmol–1 K–1 and the molecular weight of hydrogen = 2,
calculate (i) mass of one molecule of hydrogen and (ii) Boltzmann constant.
Solution
(i) Mass of one molecule of H2
Molecular weight
= _______________
Avogadro No.
2
= __________ g = 3.32 × 10–24 g
6.02 × 1023
= 3.32 × 10–27 kg

(ii) Boltzmann constant


R 8.31
k = __ = __________
N 6.02 × 1023
= 1.38 × 10–23 Jmol–1 K–1
Example 2.3 At what temperature, the root mean square speed of molecules of a
gas will be double the root mean square speed at 0°C.
Solution Given T1 = 0°C = 273 K
v2 rms = 2v1 rms
____
3RT
vrms = ____
Since
÷M
___
v2 rms T2
\ _____
÷__
v1 rms = T1
Substituting the values
____
T2
÷
2 = ____
273
or T2 = 4 × 273 = 1092 K
or T2 = (1092 – 273)°C = 819°C
Example 2.4 The density of CO2 gas at 0°C and 105 Nm–2 is 1.98 kg m–3. If
pressure remains constant, calculate the root mean square speed of its molecules
at 0°C and then, at 30°C.
5 –2 –3
Solution Given P = 10 Nm , r = 1.98 kgm
At 0°C, the root mean square speed of molecules is
___ _______
3 × 105
3P
÷r 1.98÷
vrms = ___ = _______ = 3.89 × 102 ms–1
Kinetic Theory 2.45

The r.m.s. speed is directly proportional to the square root of absolute


temperature of gas i.e.,
__
vrms μ ÷T
___
v1 rms T
\ v2 rms ÷
____ = __1
T2
Here T2 = 30°C = (273 + 30) K = 303 K

T2 = 0°C = 273 K
___
T2
\
÷
v2 rms = v1 rms × __
T1
____
303
____
or v2 rms = 3.89 × 10 2
÷273
= 4.1 × 10 ms–1
2

Example 2.5 Calculate the translational kinetic energy of 1 mole gas at 0°C.
Solution Number of degrees of freedom per molecule for translational motion
= 3.
Accordingly, mean kinetic energy of a molecule
1
= 3 × __ kT
2
3
= __ kT
2
\ Mean translational kinetic energy per mole
3
= NA × __ kT
2
3
= __ RT
2
3
= __ × 8.3 × 273 = 3.4 × 103 J
2
Example 2.6 Calculate the temperature at which the root mean square velocity of
a H2 molecule is 3/2 times that of a O2 molecule at 27°C. The molecules weight
of H2 is 2 and that of O2 is 32.
Solution Let T1 be the temperature of H2 molecule of molecular weight m1. Then
the r.m.s. velocity of H2 molecule is
____
3kT1
÷
C1 = ____
m1
2.46 Heat and Thermodynamics

If T2 be the temperature of O2 molecule of molecular weight m2, then its r.m.s.


velocity is
_____
3kT2
÷
C2 = ____
m2
3
We have C1 = __ × C2
2

or ( )
3kT1
____
m1
1/2
3 3kT2
( )
= __ ____
2 m2
1/2

T1 __
___ 9 ___T2
or m1 = 4 × m2
9 m1
or T1 = __ ___ T
4 m2 2
9 2
= __ × ___ × 300 K
4 32
= 42.19 K
Hence, the temperature of H2 gas in °C is
= 42.19 – 273
= – 230.81°C
Example 2.7 The r.m.s. speed of the smoke particles suspended in air at 200 K is
2.8 cm/s. The mass of each smoke particle is 1.55 × 10–17 kg.
Calculate the Avogadro’s number, taking R = 8.31 J/g mole K.
Solution Given T = 300 K,
m = 1.55 × 10–17 kg
vrms = 2.8 cm/s = 0.028 m/s
R = 8.31 J/g mole K
The smoke particles in air behave like the molecules of a perfect gas and their
average kinetic energy at temperature T will be 3/2 kT.
1
__ 3
\ m(vrms)2 = __ kT
2 2
m(vrms)2
________
or k=
3T
1.55 × 10–17 × (0.028)2
= ___________________
3 × 300
= 1.35 × 10–23 J/K
Kinetic Theory 2.47

Avogadro’s number
R
NA = __
k
8.31
= __________
1.35 × 10–23
= 6.16 × 1023 per gm-mole
Example 2.8 Calculate the mean free path and collision frequency of air under
normal condition when the radius of the air molecule is 10–10 m, number density
is 1025 m–3 and mean velocity is 500 m/s.
Solution The mean free path is given by

1
l = _______
__
÷2 pd2n
and the collision frequency
_
c
n = __
l
_
where n is the number density, d is the diameter of air molecule and c is the mean
velocity. Here
n = 1025 m–3
d = 2 × 10–10 m
_
c = 500 m/s
Then
1
l = _____________________
__
÷2 × p (2 × 10–10)2 × 1025

= 5.63 × 10–7 m
and
500
n = __________
5.63 × 10–7
= 8.89 × 108 s–1
Example 2.9 Calculate the probability that a molecule of helium at temperature
0°C under the pressure 50 Pa covers the distance 0.75 mm without collisions: The
diameter of helium molecule is 2.6 × 10–8 cm.
Solution The probability of covering a distance x without collisions is

x
[ ]
f(x) = exp – __
l
2.48 Heat and Thermodynamics

where l is the mean free path, given by


1 kT
l = _______
__ = ________
__
÷2 p d n ÷2 p d 2P
2

Here
d = 2.6 × 10–8 cm,
x = 0.75 mm = 0.075 cm
k = 1.38 × 10–16 erg/K
T = 273 + 0 = 273 K
P = 50 Pa = 50 × 10 dynes/cm2
and \
1.38 × 10–16 × 273
l = ____________________
__
÷2 p (2.6 × 10–8)2 × 500

= 0.0251 cm.
Therefore, probability

[0.075
f(0.075) = exp – ______
0.0251 ]
= 5 × 10–2
Example 2.10 Using mean free path concept, show that in a gas at ordinary
pressures, the number of molecules striking unit area of the wall per second is
_
1/4 nc.
Solution If Pc be the collision probability, then the number of collisions suffered
by the molecules having velocity between c and c + dc per unit volume per second
is 1/2 Pc nc dc where nc dc is the number of molecules per unit volume in that
velocity range. The factor 1/2 comes because each molecule is counted twice in
the collision with other molecules. The number of molecules per unit volume
coming straight from collision in 1 sec. is
c
( )
Pc nc dc = __ nc dc
lc
where lc is the mean free path of the molecules moving with velocity c.
Let us consider the molecules striking the area dA of the surface in a direction
making an angle q with the normal to the surface contained within the solid angle
dw.
Coming from a distance lying between x and x + dx in the above mentioned
direction, these molecules will be contained in an element of volume x2 dx dw.
Kinetic Theory 2.49

The total number of molecules in this volume


c
= __ nc dc x2 dx dw.
lc
dA cos q
The fraction of molecules in the desired direction = ________ . Out of these
4p x2
molecules, only the fraction e– x/lc can reach the surface dA without suffering a
collision. Hence, the number of molecules striking the area dA in 1 sec. is
• p/2 •
1
___ 1
dA Ú nc dc × c Ú 2p sin q cos q dq Ú __ e– x/lc dx
0 0 4p l
0 c

1
()
= dA Ú c nc dc __ × 1.
0 4

The number striking unit area per sec.



1 1 _
= __ Ú c nc dc = __ nc.
40 4

Example 2.11 If the viscosity of nitrogen be h = 1.66 × 10–5 kg m–1 s–1, the
_
average velocity of the molecules c = 4.5 × 102 m/s and the density r = 1.25 kg/
m3 and n = 2.7 × 1025 molecules/m3 for nitrogen, calculate the mean free path,
collision frequency and the molecular diameter.
1 _
Solution h = __ r cl
3
3h
\ Mean free path l = ___
_
rc
3 × 1.66 × 10–5 kg m–1 s–1
________________________
=
1.25 kg m–3 × 4.5 × 102 m s–1
= 8.85 × 10–8 m
Collision frequency =
_
4.5 × 102 m s–1
c _____________
__ =
l 8.85 × 10–8
= 5.08 × 109 s–1
1
l = _______
__
÷2 p nd2
_______
1
d = _______
or diameter
÷
__
÷2 p nl
2.50 Heat and Thermodynamics

[ 1
= _________________________________
__
÷2 (3.14) 2.7 × 10 m × 8.85 × 10–8 m
25 –3 ] 1/2

= 0.031 × 10–8 m
Example 2.12 Calculate the viscosity coefficient of H2 gas at NTP from the
following data
The density r = 8.96 × 10–2 kg.m–3
The mean free path l = 1.69 × 10–7 m,
The Boltzmann constant
k = 1.38 × 10–23 J/K.
Solution Coefficient of viscosity
1 _
h = __ rcl
3
1 _
= __ × 8.96 × 10–2 kg.m–3 × c × 1.69 × 10–7 m
3
____
_ 8kT
Now pm÷
c = ____

Mass of hydrogen molecule


2.0 gm
= ___________
6.023 × 10–23
= 3.32 × 10–27 kg

[ ]
1/2
_ 8 × 1.38 × 10–23 J K–1 × 273 K
\ c = _________________________
3.14 × 3.32 × 10–27 kg
= 48.09 × 102 ms–1
1
\ h = __ (8.96 × 48.09 × 1.69) kg m–1 s–1
3
= 242.7 × 10–7 kg m–1 s–1
= 2.43 × 10–5 kg m–1 s–1
Example 2.13 Transform the Maxwell distribution function going from the
variable v to the variable u = v/vmp where vmp is the most probable speed of the
molecules.
Solution The Maxwell’s distribution function is
__
f(v) = __
m
2 ___
÷ ( )
p kT ( )
mv2
v exp – ____
3/2 2
2kT
Kinetic Theory 2.51

We know that
____
2kT
vmp = ____
m ÷
Consequently,
__
2
f(v) = __
p ÷
× 2
__ m
÷2 ____
2kT ( ) 3/2 2
v [ ( )] mv2
exp – ____
2kT

÷p v3
v2
4__ ____
= ___
mp
exp
( )
– v2
____
v2mp

or
dnv ___
____ 4 v2
= __ ____
ndv ÷p v3mp
exp
– v2
____
v2mp ( )
\

v
÷p v3
v2
4__ ____
dn = n × ___
mp
exp
– v2
____
v2mp ( )
dv

Since ___
v ∫u
mp
\ dv = vmp du
Therefore,

÷p v3
v2
4__ ____
dn = n × ___
mp
exp
– v2
____
v2mp
du
( )
4__ 2
or dn = ___ nu exp (– u2) du
÷p

Example 2.14 Calculate the value of velocity component vx corresponding to


which the number of molecules is 1/2 the number of molecules corresponding to
the maximum probable velocity.
Solution The Maxwell’s velocity distribution law is
2
mv
m 1/2 – ____
x
N(vx) = N _____
2pkT (e 2kT )
The number of molecules corresponding to maximum probable velocity (vx = 0)
is
m 1/2
|N(vx)|max = N _____(
2pkT )
1
\ For N(vx) = __ |N(vx)|max, we have
2
mvx2
m 1/2 – ____ 1 m
(
N _____
2pkT ) e 2kT (
= __ N _____
2 2pkT ) 1/2
2.52 Heat and Thermodynamics

mvx2
– ____ 1
or e 2kT = __
2
2
or e mvx /(2kT) = 2

mvx2
or ( )
____ = log 2 = 0.693
2kT e

\ (
1.386 kT
vx = ± ________
m ) 1/2

QUESTIONS
1. What is an ideal gas? Derive an expression for pressure exerted by an ideal
gas. Mention the basic assumptions.
2. Give the fundamental assumptions of the kinetic theory of gases. Show
that the pressure exerted by a perfect gas is two-third the kinetic energy of
molecules per unit volume.
3. What are the essential features of the kinetic theory of gases. Apply it to
explain the Boyle’s law, Charle’s law and Avogadro’s law.
4. Derive Maxwell’s law of distribution of velocities for the molecules of a
gas. Give a suitable experiment for its verification.
5. On the basis of kinetic theory of gases, deduce an expression for
(a) the mean velocity
(b) the most probable velocity
(c) the root mean square velocity
6. State and prove the classical law of equipartition of energy. Discuss its
importance.
7. Show that the number of molecules is the speed range c and c + dc is

m
N(c) dc = 4pN _____
2pkT ( )
3/2
e–mc /2kT c2 dc
2

8. Write the Maxwell-Boltzmann law of distribution of molecular speeds of


a gas. Use it to find
(i) maximum probable speed
(ii) average speed and
(iii) root mean square speed
9. On the basis of kinetic theory of gases, prove that the r.m.s. speed of the
molecules of a gas is given as
Kinetic Theory 2.53

___
3P
vrms = ___
÷ r
where P is the pressure and r the density of gas.
10. What is Brownian motion? Explain it. On the basis of Einstein’s theory of
Brownian motion, deduce an expression for Avogadro’s number.
11. Prove that the mean energy corresponding to each square term in the
expression for energy of a system (or corresponding to each degree of
freedom) in thermal equilibrium at absolute temperature T is 1/2 kT where
k is Boltzmann constant. Then prove that the average kinetic energy of a
gas molecule is 3/2 kT.

OBJECTIVE TYPE QUESTIONS


1. According to kinetic theory of gases
(a) the rate of diffusion of gas is directly proportional to the root mean
square spped of molecules.
(b) mean kinetic energy per molecule is directly proportional to the square
root of absolute temperature of the gas.
(c) root mean square speed of molecules is directly proportional to its
pressure.
(d) pressure of gas is directly proportional to the mean speed of
molecules.
2. Brownian motions are
(a) regular (b) random
(c) irregular and random (d) continuous
3. Verification of the kinetic theory is given by
(a) circular motion (b) Brownian motion
(c) both (a) and (b) (d) none of these
4. The pressure of gas P and its average kinetic energy per unit volume E are
related as
E
(a) P = E (b) P = __
2
2 3
(c) P = __ E (d) P = __ E
3 2
5. The expression for root mean square speed of the molecules of a gas is
__ ___
÷P 3P
(a) ___
r
(b)
÷___r
2.54 Heat and Thermodynamics
__
÷P ____
(c) ___ (d) ÷5Pr
3r
6. The ratio of root mean square speeds of two gases of molecular weights m1
and m2 at a temperature T will be
___
m1 m1
(a) ___ ___
m2 (b)
÷ m2
___
m2 m
(c) ___ (d) ___
÷
2
m1 m 1
7. If a gas is at a temperature T°K, the root mean square speed of its molecules
will be proportional to
__ 1__
(a) ÷T (b) ___
÷T
(c) T2 (d) T
8. The number of degrees of freedom of a diatomic gas molecule are
(a) 6 (b) 7
(c) 5 (d) 3
9. The mean total kinetic energy of one mole of a gas is
1
(a) kT (b) __ kT
2
3 3
(c) __ kT (d) __ RT
2 2
10. The number of degrees of freedom of a mono atomic molecule is
(a) 3 (b) 4
(c) 5 (d) 6
11. The mean free path of molecules is
(a) the mean distance between the molecules
(b) the free distance of molecules
(c) the distance travelled by the molecules
(d) the average value of distance travelled by the molecules between two
successive collisions
12. The phenomenon produced due to transfer of momentum is
(a) thermal conductivity (b) viscosity
(c) diffusion (d) none of these
13. The expression for mean free path of molecules of any gas is
1 1
(a) l = ________
___ (b) l = ______2
÷2p n d2 p nv d
___
(c) l = ÷2p n d2 (d) p r d2
Kinetic Theory 2.55

14. The viscosity in a gas arises due to


(a) transport of temperature (b) transport of molecules
(c) transport of momentum (d) transport of energy
15. The coefficient of viscosity of a gas is proportional to
1__ 1
(a) ___ (b) __
÷T T
(c) T1/2 (d) T
16. The relation between the coefficient of viscosity h and coefficient of
thermal conductivity K of any gas is
h
(a) K = hcv (b) K = __
r
(c) K = h/cv (d) K = hr

Answers
1. (a); 2. (c); 3. (b); 4. (c); 5. (b); 6. (c); 7. (a);
8. (c); 9. (d); 10. (a); 11. (d); 12. (a); 13. (b); 14. (c);
15. (c); 16. (a).
Real Gases, Liquefaction of Gases

3
Chapter
and Production and Measurement
of very Low Temperatures

3.1 PERFECT GAS AND REAL GAS


The p, v, Q relationship for gases, liquids and solids can be expressed in the form
of equation of state of the general type:
f(p, v, Q) = 0 (1)
Only in the case of gases, much progress has been made in the development of
such equations. The equation of state for the perfect gas is
pv = RQ (2)
For a real gas, the equation of state is not so simple. Instead of representing
our measurements in the form of an equation of state, we could only present the
relationship between p, v, Q in the form of a table of experimental numbers or a
set of curves as in Fig. 3.1.

1.1

0°K
H 2 30
H 2 600°K
Ideal gas
1.0

pv
nRQ
H2

0.9
O
CO 2

60

K
300
°K

0.8
0 20 40 60 80 100
p

Fig. 3.1 Curves representing equations of state for a few real gases [After: B.K. Agarwal
Thermal Physics, Lokbharti Publications Allahabad India (1988)] the figure shows
departure from the perfect gas law
3.2 Heat and Thermodynamics

3.2 EXPERIMENTS ON THE BEHAVIOUR OF REAL GASES


A perfect (or ideal) gas when compressed isothermally (i) obeys Boyle’s law and
(ii) remains a gas no matter how great a pressure is applied to it. All real gases
show departure from Boyle’s law, particularly at high pressures, and also become
liquids unde suitable (high) pressure and (low) temperature conditions.

(a) Amagat’s Experiments on Departure from Boyle’s Law


The so-called permanent gases (for example, air, oxygen, nitrogen and hydrogen)
obey Boyle’s law to within less than one part in a thousand at ordinary pressures
and temperatures. At higher pressures and lower temperatures the departures
become pronounced.
Regnault in 1847 performed a series of experiments to verify the law. He
increased the pressure range upto 27 atm. He found that hydrogen, nitrogen, air
and carbon dioxide did not obey Boyle’s law strictly.
Amagat, working in France around 1878, increased the pressure upto 400 atm.
For this he pumped mercury in the manometer of the apparatus, used for the
verification of Boyle’s law, with the mercury reservoir fixed to the face of a high
cliff in one experiment and to the side of a coal mine shaft in another.
Amagat performed his famous experiment on nitrogen in 1893 in which the
apparatus was set up at the bottom of a coal mine. It consisted of a large cylinder
B of mercury provided with a large screw-plunger D at one end, a small calibrated
tube C containing nitrogen in the middle, and a steel manometer tube A about
1000 feet long extending up the mine shaft at the other end, Fig. 3.2. The gas in
the small tube C could be maintained at the desired temperature by means of a
suitable oil bath.

Fig. 3.2 Amagat’s set-up for nitrogen


Real Gases, Liquefaction of Gases and Production and... 3.3

The result of these experiments showed that no gas is perfect, that is, no gas
obeys Boyle’s law. The nature of deviations can be displayed by plotting pv against
p, Fig. 3.3. Experiments of this kind were also performed later on by Holborn and
Otto in Berlin and by Kamerlingh-Oanes and Keesom in Leiden. These days such
measurements are made with modern high pressure techniques at various bureaus
of standards. It has been shown that the inert gases give curves similar to that of
hydrogen. Gases which are most easily liquefied (for example, CO2) deviate the
most from the perfect gas behaviour.

N2

H2
pv

RQ 2
RQ 3
Perfect gas
CO 2

Fig. 3.3 pv vs. p plot

It is seen that from Fig. 3.3 that as p Æ 0, the product pv approaches the same
value for all gases (including perfect gas) at the same temperature,

lim (pv) = function of temperature only, independent of gases


pÆ0

= RQ (3)

lim (pV) = nRQ (4)


pÆ0

(b) Andrew’s Experiments on Carbon Dioxide


Another difference between a perfect gas and a real gas in an isothermal
compression is shown in Fig. 3.4. Suppose the gases are compressed as a given
suitable temperature in two similar cylinders each provided with a piston and a
pressure gauge. When the piston in Fig. 3.4a is pushed to the right, the pressure
reading in the gauge rises steadily and the p-V graph is the familiar Boyle’s law.
On the other hand, for the real gas the pressure rises at first along the curve AB,
Fig. 3.4b, in a manner not very different from the case (a). Then a sharp break
occurs at the point B accompanied by the appearance of liquid drops on the walls
3.4 Heat and Thermodynamics

C Perfect gas Real gas


F

P B P
D C B
E

A A

v v

VA VA
pA pA

VB VB
pB pB
VC VC

pC pC = pB
VD
pD = pB
VE
pE = pB

VF
pF
(a) (b)

Fig. 3.4 Isothermal compression of (a) a perfect gas, and (b) a real gas

of the cylinder. As the piston is further pushed to the right, the volume continues
to decrease without any increase of pressure (curve BCDE parallel to V-axis).
This is the process of liquefaction or condensation, beginning at B and ending at
E. As liquids are nearly incompressible, beyond the point E considerable increase
in the pressure has to be employed to achieve any reduction in volume. Therefore,
the curve rises nearly vertically from E to F.
Such isothermal curves at several temperatures were obtained by Thomas
Andrews in 1863 for carbon dioxide with a simple apparatus shown in Fig. 3.5.
Carbon dioxide was enclosed in a calibrated glass tube by a pellet of mercury,
the open end of the tube being in a water chamber to which pressure was applied
by screwing in a plunger. The pressure was measured by noting the decrease in
volume of air in a similar tube. The portions of tubes containing CO2 and air were
enclosed in two separate water baths, with the one around CO2 maintained at any
desired temperature between 0°C and 100°C and the other around air maintained
at a constant temperature throughout.
Real Gases, Liquefaction of Gases and Production and... 3.5

............................... Temperature ............................


............................
Heater ...............................
...............................
...............................
............................
............................
............................
............................... ............................
...............................
...............................
...............................
bath ............................
............................
...............................
............................... ............................
............................
............................... ............................
...............................
...............................
............................... CO2 Air ............................
............................
............................
...............................
............................... ............................
............................
...............................
............................... ............................
............................

Constant temperature bath


...............................
............................... ............................
...............................
............................... ............................
............................
...............................
............................... ............................
............................
Mercury pellet ............................

Water

Fig. 3.5 Andrews’ apparatus

The resulting curves are shown in Fig. 3.6. These curves can be divided into
two classes:
1. At 31.1°C and above: The p-v curves are similar to though not so steep as,
the curves for air.
2. Below 30°C: Each portion shows three distinct portions, as in Fig. 3.4b.
For a typical curve abcd at 21.5°C the CO2 is gaseous (vapour) between a
and b, liquefying between b and c (both liquid and vapour being present
together in equilibrium), and liquid beyond c.

48.1°C
100

90
p atm

80 P S O
31.1°C
d
x Gas
Liquid

70 N
Critical
c 21.5°C b isotherm
60 Liquid + Va
vapour M pour
50
13.1°C Q
0 1 2 3 4 5 6 7
V CC

Fig. 3.6 Andrews’ curves for carbon dioxide


3.6 Heat and Thermodynamics

The pressure corresponding to part bc is called the vapour pressure of the


liquid. In the tube the liquid will appear more dense than the vapour, with a
meniscus between them. As the temperature is raised, the length of the horizontal
portion decreases (points b and c move towards each other), showing that the
difference between the nature of the vapour and that of the liquid is less. That
is, the density of the liquid decreases while that of vapour increases. Just below
31.1°C the horizontal part is reduced to a point (points b and c coincide at X).
At this temperature the densities of liquid and vapour have become identical and
the meniscus separating liquid from vapour disappears. At 32.5°C (not shown in
Fig. 3.6) there is no horizontal part at all, though there is a kink in the curve which
disappears as the temperature is further raised. At 48.1°C the curve is smooth and
very similar to the air curves.

3.3 CRITICAL POINT


The temperature at which the meniscus (or equivalently the horizontal part of the
isotherm) just disappears is called the critical temperature Qc of the substance.
The vapour pressure at Qc is called the critical pressure pc. The specific volume
at Qc and pc is called the critical specific volume vc. These critical constants for
some of the substances are given in Table 3.1.

Table 3.1 Critical constants for real gases.


Qc
Substance pc (atm) vc (cc/mole)
°K °C
Helium 5.3 –268 2.26 57.6
Hydrogen 33.3 –240 12.8 65.0
Nitrogen 126.1 – 147 33.5 90.0
Oxygen 154.4 – 119 49.7 74.4
Carbon dioxide 304.2 31 73.0 95.7
Water 647.2 374 217.7 45.0

It is usual to refer to a gas in equilibrium with its liquid (i.e., a saturated vapour)
or to a gas below its critical temperature as a vapour. The properties of vapour
differ in no essential way from those of a gas.
It is clear from the curves of Andrews that Qc is the temperature above which
the substance cannot be liquefied, however, great the pressure. The point X with
coordinates pc, vc on the isotherm for the critical temperature Qc is called the
critical point. At the critical point the substance passes from the gaseous to the
liquid state.
Real Gases, Liquefaction of Gases and Production and... 3.7

3.4 CONTINUITY OF STATE


At the point M on the isotherm abcd (Fig. 3.6), the substance is in the vapour
phase and at the point P it is in the liquid phase. When we go from M to P along
abcd the discontinuity represented by horizontal part bc of the curve occurs.
It is possible to go from M to P without any discontinuity occurring if we
avoid the region bounded by the dotted curve bXc. For example, we can follow
the path MNOP. The vapour at M, at a temperature below Qc, is heated at constant
volume to point N, above Qc. It is now compressed along the isotherm to point
O, and finally cooled at constant pressure so that the volume decreases along the
dotted line OP. The substance which was in the vapour phase at M is now in the
liquid phase at P on the same isotherm. At no point along the path MNOP were
the two phases (liquid and vapour) simultaneously present. The transition occurs
at the point S on the critical isotherm. To its right the system is in gas state and
to the left in liquid state. Thus, the transformation from vapour to liquid occurs
smoothly and continuously. Above pc it is possible simply by cooling, to pass from
the gaseous state to the liquid state without any mixture of phases. This is called
the continuity of the liquid and gaseous states.
James Thomson suggested that the horizontal part
of the isotherm like bc in Fig. 3.6, can be replaced d

by a curve ablmncd with continuously changing


P
slopes, as shown in Fig. 3.7. Such a continuous
l
curve is typical of a cubic equation with three real b
c m
roots. A short portion of the curve above b can be a
n
realized in practice by supersaturated vapour, and a
small part below c by superheated liquid. However, v
these are metastable states and the rest of the dotted Fig. 3.7 Continuous curve
curve is not observed.
The importance of the suggestion lies in the fact that the whole family of
isotherms for real gas can now be represented by a simple single equation of state.
This would be difficult for the discontinuous curve abcd.

3.5 FORM OF EQUATION OF STATE


Kamerlimgh-Onnes in 1901 represented the experimental results by expressing
pv as a power series (or virial expansion) in p or 1/v,

( 1
pv = A 1 + B __
1
__
v + C v2 + … ) (5)

at a constant temperature. Here A, B, C, etc. are functions of temperature and are


called the virial coefficients. A is called the first virial coefficient. Clearly for a
perfect gas A = RQ and all other virial coefficients are zero.
3.8 Heat and Thermodynamics

Experimental data for a real gas can be fitted closely if large number of virial
coefficients are included in the expansion. However, soon the computation problem
becomes difficult. Therefore, it is useful to consider simple modifications of the
perfect gas equation.
A Dutch physicist van der Waals suggested a useful form for the real gases in
1873. His equation is

( p + ___Va ) (v – b) = Rq
2
(6)

where a and b are constants. The importance of this equation lies in its general
application as an approximate equation over wide ranges of variables and for
a wide range of gases. For a gas occupying a large volume, the van der Waals
equation reduces to the perfect gas equation.

3.6 VAN DER WAALS EQUATION AND THE CRITICAL POINT


Figure 3.8 shows isotherms plotted according to the van der Waals equation

Rq a
p = ______ – __2 (7)
(v – b) V
or pv3 – (Rq + bp)v2 + av – ab = 0.
The van der Waals equation is cubic in v. We obtain for q < qc (the critical
temperature), either three real roots (three values of v for a single value of p as at
x, y, and z in Fig. 3.8) over a certain range of values of p, or two imaginary roots
and one real root at pressures above and below this range. As the temperature
increases, the three real roots approach one another and at qc, they become equal

Fig. 3.8 Isotherms for a van der Waals gas


Real Gases, Liquefaction of Gases and Production and... 3.9

(critical point X). The maxima and minima points (like A and B) also merge into
one point X, which is, therefore, the point of inflexion with a horizontal tangent.
At q > qc, only one real root exists for all values of p.
The maxima and minima of the isotherms for q > qc are not usually observed.
However, by using very pure gases and liquids, the portions xA (supersaturated
vapour) and By (supercooled liquid) can be experimentally realized. They are
metastble states.
We have drawn the locus of the maxima and minima of the isotherms as a
dotted curve in Fig. 3.8. Its equation is found by the condition that the tangent to
dp
( )
the isotherm is horizontal, ___ = 0 where q = constant i.e., from Eqn. (7)
dv q

( )
dp
___
dv
Rq 2a
= – _______2 + ___3 = 0
q
(v – b) v
(8)

The maximum of the dotted curve determines the critical point X. It is obtained
by the condition that the tangent to the dotted curve described by Eqn. (8) is
horizontal,

( )
dp
d ___
___
dv dv q ( )
d 2p
∫ ___2
dv q
=0 (9)

2Rq
_______ 6a
3
– ___4 = 0 (10)
(v – b) v
At the critical point X (pc, vc, qc), the three Eqns. (7), (8) and (10) must be
simultaneously satisfied i.e.,
Rqc a
pc = _______ – ___2 (11)
(vc – b) vc
– Rqc
_______ 2a
2
+ ___3 = 0 (12)
(vc – b) vc
2Rqc
_______ 6a
3
– ___4 = 0 (13)
(vc – b) vc
From Eqns. (12) and (13), we get
vc _____
__ vc – b
=
3 2
or vc = 3b (14)
Substituting this value in Eqn. (12), we get
8a
qc = _____ (15)
27bR
3.10 Heat and Thermodynamics

Equation (11) gives


a
pc = ____2 (16)
27b
Thus, the constants a and b can be determined from the critical point data. It
may be verified that

Rqc __
____ 8
pc vc ∫ 3 = 2.67 (for all gases) (17)

The observed values of Rqc/(pc vc) fall in the neighbourhood of 3.5. Thus van
der Waals equation is not closely obeyed by any gas in the neighbourhood of its
critical point. However, it is still a very convenient equation over a wide range.

Values of a and b in Terms of Critical Constants


From Eqns. (15) and (16),

q2c [8a/(27Rb)]
___
2
64a
___________ = ____
pc = 2
a/(27b ) 27R2
Therefore,
2
27R2 q c
a = ____ ___ (18)
64 pc
and
qc _________
__ 8a/(27Rb) ___
8a
pc = a/(27b2) = R

R qc
\ b = __ __ (19)
8 pc

Thus, knowing qc and pc, the values of van der Waals constants a and b can be
calculated.

Gas and Vapour


In general conversation, when the substances which are solid ir liquid at normal
temperature and pressure come in gaseous state are known as vapour whereas the
substances which are in gaseous state at normal temperature and pressure are said
to be gases. In scientific language, the gaseous state above the critical temperature
of a substance is known as gas and the gaseous state below the critical temperature
of the substance is known as vapour.
Real Gases, Liquefaction of Gases and Production and... 3.11

3.7 COMPARISON BETWEEN EXPERIMENTAL


AND THEORETICAL P‐V CURVES
On comparing the experimental P-V curve (Fig. 3.9(a)] and theoretical P-V curve
obtained from van der Waals Equation of state (Fig. 3.9b), we see that Bb and cC
parts obtained in the theoretical curve (Fig. 3.9b) which respectively represent the
supersaturated vapour state and supersaturated liquid state, are not obtained in the
experimental curve (Fig. 3.9a). On the other hand, only a horizontal line BC is
obtained in the experimental curve which shows that the gas always remains in
steady state. Indeed, the supersaturated vapour state (in which volume decreases
with the decrease in pressure) and supersaturated liquid state are the unsteady
states which are impossible to be obtained in practice.

Fig. 3.9 P-V isotherms for van der Waals equation of state (a) obtained from
Andrew’s experiment for CO2; (b) theoretical curves for CO2

Example 3.1 Show that the van der Waals gas law departs from perfect gas law
by 62.5% at critical temperature.
Solution The van der Waals equation is

( p + ___Va ) (V – b) = RT
2

a ab
or pV – bp + __ – ___2 = RT (i)
V V
The perfect gas equation is
pV = RT (ii)
The deviation is obtained by subtracting Eqn. (ii) from Eqn. (i) at Tc:
a ab
– bp + __ – ___2 = R (Tc – T)
V V
3.12 Heat and Thermodynamics

a ab
or RDT = – bp + __ – ___2
V V
8a a
Now, Tc = _____, pc = ____2 , Vc = 3b
27Rb 27b
So, at the critical point
a a ab
RDT = – b ____2 + ___ – ___2
27b 3b 9b
Using Tc value i.e.,
27RTc
a ______
__ = ,
b 8
we have
1 27RTc 1 27RTc 1 27RTc
RDT = – ___ × ______ + __ × ______ – __ × ______
27 8 3 8 9 8
DT __
___ 5
\ =
Tc 8
DT __
___ 5
or = × 100%
Tc 8
= 62.5%

3.8 VAN DER WAALS EQUATION


AND THE BOYLE TEMPERATURE
In general, instead of the Boyle law, pv = constant a fixed Q, the real (or imperfect)
gases obey an empirical power equation of the form
pv = A + Bp + Cp2 + Dp3 + … (1)
where A, B, C, D, … are the virial coefficients. Usually C, D, … are very small.
Therefore, apart from A, the most important coefficient is B. The variation of pv
and p is shown in Fig. 3.10.

Fig. 3.10 Variation of pv with p for a real gas at different temperatures


Real Gases, Liquefaction of Gases and Production and... 3.13

If we find the slope of the curves near p = 0, we have

( )
∂(pv)
_____
∂p p=0
= (B + 2Cp + 3Dp2 + …)p = 0

=B (2)
We define a temperature QB at which B = 0. From Fig. 3.10, we see that B is
negative at Q < QB and positive at Q > QB. For Q = QB and p Æ 0 we have B = 0
so that the expansion (1) reduces to
pv = A
that is, Boyle law holds approximately again. For this reason QB is called the
Boyle temperature.
For a van der Waals gas we can find QB by calculating ∂(pv)/∂p and equating it
RQ a
to zero at p = 0. Thus van der Waals equation pv = _____ – __2 gives
v–b v
RQv a
pv = _____ – __
v–b v
∂(pv)
_____
∂p [1 v a
= RQ _____ – RQ ______2 + __2
v–b (v – b) v ]( )∂v
___
∂p Q

[ RQb
(v – b)
a
= – _______2 + __2
v ]( ) ∂v
___
∂p Q
(3)

If we require it to be zero at Q = QB and p = 0, it is equivalent to considering v


as approaching infinity so that v – b can be replaced by v. Therefore, from Eq. (3),
we have
– RQB b + a = 0

a 27
or QB = ___ = ___ Qc (4)
Rb 8
The van der Waals equation can be written as

(
b
pv = RQ 1 – __
v ) –1
a
– __
v
b b__2
(
= RQ 1 + __
a
__
v + v2 + … – v )
[
= RQ 1 + _________
v
b2
b – (a/RQ) __
+ 2+…
v ] (5)

At QB, we can put Q = QB = a/Rb. Then for p small, or large v, Eqn. (5)
reduces to
3.14 Heat and Thermodynamics

pv = RQB (6)
for a real van der Waals gas.
It is useful to plot (pv/RQ) vs. p as shown in Fig. 3.11 for CO2. Then all the
curves converge exactly to the same point on the vertical axis, whatever the
temperature, because in the limit of low pressure lim (pv) = A = RQ. The curve at
QB is parallel to the p-axis for small p values.

Fig. 3.11 (pv/RQ) vs. p graph for CO2 [After B.K. Agarwal, Thermal Physics, Lokbharti
Publications, Allahabad, India (1988)]

3.9 CORRESPONDING STATES


The van der Waals equation can be put in a form that is applicable to any substance
by introducing the reduced quantities pr, vr, Qr as follows
p v
pr = __
p, vr = __
v,
c c

Q
Qr = ___ (7)
Qc
Then,
a
p = ____2 pr, v = 3b vr,
27b
8a
Q = _____ Qr (8)
27Rb
and the van der Waals equation becomes

( 3
)
pr + ___2 (3vr – 1) = 8Qr
vr
(9)
Real Gases, Liquefaction of Gases and Production and... 3.15

This is a remarkable equation because it does not involve the constants a and b
characteristic of a particular gas. It holds for any van der Waals gas. The statement
that it is a universal equation valid for all gases is called the law of corresponding
states to the extent that real gases obey the van der Waals equation. Different
gases with the same pr, vr and Qr are said to be in corresponding states.

3.10 JOULE’S EXPANSION OF AN IDEAL GAS


AND VAN DER WAAL’S GAS
When a definite mass of a gas is expanded such that the external work done by
the gas or on the gas is zero, neither the gas absorbs nor rejects any heat, then the
expansion of gas is known as free expansion or Joule’s expansion. Scientist Joule
in his experiments, studied the expansion of ideal gas and real (or van der Waal)
gas and concluded that there exists no force between the molecules of an ideal gas
whereas there exist attractive forces between the molecules of van der Waal gas
which are known as van der Waal forces.
Experimental Arrangement Figure 3.12
shows the experimental arrangement of
Joule’s experiment. It consists of two copper
cylindrical vessels A and B which are mutually
connected by a tube C. There is a stop cock
S in the tube C. Vessel A is filled with the
experimental gas at a high pressure while the
vessel B is kept completely evacuated. Both
the cylinders are immersed in a vessel filled
with water. Fig. 3.12 Experimental arrangement
Temperature of water is measured with of Joule’s experiment
the help of a sensitive thermometer T. It is
clear that on opening the stop cock S, free expansion of experimental gas from
cylinder A to cylinder B takes place since it involves neither any external work
done nor any exchange of heat.

(a) Joule’s Expansion of an Ideal Gas


Assuming air to be an ideal gas when scientist Joule filled air in the cylinder A
and then allowed the air to expand freely by opening the stop cock S, he found
that there is no change in temperature of water. It is thus concluded that there
are no intermolecular attractive forces acting between the molecules of an ideal
gas because if there will be any attractive force between the molecules, then the
temperature of water should fall during this expansion because there should be
some internal work done by the gas against the attractive forces in separating the
molecules apart from each other.
3.16 Heat and Thermodynamics

Thus, on bringing any change in pressure and volume of an ideal gas, its
temperature remains constant i.e., its internal energy remains unchanged. This is
known as Joule’s law. In other words, we can say that according to Joule’s law,
the internal energy U of an ideal gas only depends on its temperature T and does
not depend on its volume V and pressure P. Mathematically, for an ideal gas
U = f(T)

( )
∂U
___
∂V T
( )= 0,
∂U
___
∂P T
=0

and ( ) ( )
∂U
___
∂T P
∂U
= ___
∂T V
(1)

Thus, internal energy of an ideal gas remains unchanged with pressure and
volume at a constant temperature and change in internal energy per degree change
in temperature remains same whether pressure remains constant or volume
remains constant.
Remember that Joule’s law is true only for an ideal gas.

(b) Joule’s Expansion of van der Waal’s or Real Gas


When scientist Joule used a real gas at a high pressure in cylinder A in his
experiment and opened the stop cock to allow the free expansion of gas, then he
found that there is decrease in temperature of water.
From this, it is concluded that there exist inter-molecular attractive forces in
the molecules of a real gas. Hence, in expansion of gas, there is some internal
work done by the molecules against the attractive forces in going away from each
other. This increases the potential energy of gas. Since during this expansion,
neither any external heat is taken up by the gas nor the gas rejects any heat,
therefore, as a result of increase in potential energy, the internal kinetic energy
of the gas decreases. As a result, the temperature of the gas falls. It is clear that
inter-molecular forces in a real gas are of attractive nature. Remember that if
inter-molecular forces are of repulsive nature, then in free expansion of gas, the
potential energy of gas should decrease i.e., kinetic energy should increase, as a
result the temperature of gas should increase. But experimentally, it is found that
there is always cooling in Joule’s expansion of a real gas, therefore, the van der
Waal’s forces are of attractive nature only.

3.11 JOULE’S COEFFICIENT FOR REAL GAS


We have read that cooling is produced in Joule’s (or free) expansion of a real gas
(i.e., the temperature of gas falls). The temperature change in free expansion of
gas can be represented in terms of Joule’s coefficient h. If in free expansion, there
is a temperature change dT with change in volume dV, then
Real Gases, Liquefaction of Gases and Production and... 3.17

Joule’s coefficient
∂T
∂V( )
h = ___
U
(2)

Since the internal pressure caused due to inter-molecular attractive forces in a


van der Waal gas is p = a/V2 where a is van der Waal’s constant of gas and V is the
initial volume of gas, hence, the internal work done by gas against this internal
pressure in Joule’s expansion of gas is
a
dW = pdV = ___2 dV
V
Clearly, this internal work done by gas will be equal to the increase in internal
potential energy of gas or decrease in internal kinetic energy of gas. Hence,
a
Decrease in internal kinetic energy of gas = – ___2 dV (3)
V
Now, if the molar specific heat of gas at constant volume is CV and decrease in
temperature of gas is dT, then for 1 mole of gas
Decrease in internal kinetic energy = CV dT (4)
From Eqns. (3) and (4)
a
CV dT = – ___2 dV
V
dT
___ 1 a
or = – ___ ___2
dV CV V
Hence, Joule’s coefficient

∂T
∂V( )
h = ___
U
1 a
= – ___ ___2
CV V
(5)

From above Eqn. (5), since a, CV and V are the positive quantities, hence Joule’s
coefficient for a real gas is negative. It means that cooling effect is produced on
Joule’s expansion of a real gas which is in arrangement to the experimental fact.

3.12 JOULE‐THOMSON COOLING


To establish completely the existence of inter-molecular forces in real gases,
scientist Joule along with Thomson performed an experiment in 1952 which is
known as porous plug experiment. In their experiment, they found that when a
gas is passed through an insulated porous plug from a to a constant low pressure
region, then there is a change in temperature of gas. This phenomenon is known
as Joule-Thomson’s effect and this process is known as throttling process.
3.18 Heat and Thermodynamics

Porous Plug Experiment The apparatus used in the experiment is shown in Fig. 3.13.
In this experiment, experimental gas is taken in a cylinder A, compressed at a high
pressure by means of a piston P and is passed through a spiral tube S. The spiral
tube S is kept immersed in a water bath W maintained at a constant temperature so
that the temperature of the compressed gas remain constant. The compressed gas
at a constant temperature coming from the tube S enters a wide tube B, the upper
end of which is open to atmosphere. Inside the tube B, there is a porous plug G
made of cotton or silk and placed in between the two perforated discs D, D which
are surrounded by a container C filled with asbestos so that it remains insulated.
There are two platinum resistance thermometers T1 and T2 inside the tube B which
measure respectively the temperature of gas before and after passing through the
porous plug. The pressure of gas before entering the porous plug is equal to the
pressure of gas entering the tube B which is measured by a manometer (not shown
in the figure), while the pressure of gas coming out of the porous plug is equal to
the atmospheric pressure.

Fig. 3.13 Arrangement of porous plug experiment

This experiment was performed for different gases and varying the initial
pressure and temperature of the gas before entering the porous plug, different
observations were obtained. From the various observations, following conclusions
were drawn.

Results Obtained from the Experiment


In each real gas, there are intermolecular forces of the attractive nature because
on passing through the porous plug, there is a large decrease in the pressure of the
gas due to which its molecules become far separated and so, the gas has to do the
internal work against the intermolecular attractive forces. As a result, the internal
kinetic energy of the gas decreases and the gas shows the cooling effect.
Real Gases, Liquefaction of Gases and Production and... 3.19

For a given pressure difference across the plug, there is a definite temperature
for each gas, called the temperature of inversion Ti. If the initial temperature of
the gas is lower than Ti, the gas shows the cooling effect after passing through the
plug and if the initial temperature of the gas is higher than Ti, the gas shows the
heating effect after passing through the plug.

3.12.1 How Temperature Changes in Joule-Thompson Expansion


of a van der Waal’s Gas
Let 1 mole of a real gas suffer Joule-Thompson expansion at a constant pressure
P1 through a porous plug to a constant pressure P2. The initial volume and
temperature be V1 and T1 and the final volume and temperature of gas after passing
through the porous plug be V2 and T2 respectively. Refer to Fig. 3.15.
External work done on the gas by the piston 1 on left side = – P1 V1.
External work done on piston 2 by the gas on right side = P2 V2.
\ Not external work done by the gas = P2 V2 – P1 V1. (1)
Now, there are also attractive forces amongst the gas molecules and so, in this
expansion, some internal work is also done by the gas against the intermolecular
a
attractive forces. The internal pressure due to attractive forces is ___2 where a is
V
van der Waals constant of the gas. Therefore, the internal work done by 1 mole
gas in expansion of volume from V1 to V2 is
V2
a –a
V1 V
[ ]
= Ú ___2 dV = ___
V
V2
V1

a a
= ___ – ___ (2)
V1 V2
From Eqns. (1) and (2), total work done by the gas

a a
(
W = (P2 V2 – P1 V1) + ___ – ___
V1 V2 ) (3)

Now, from van der Waal’s gas equation

( P + ___Va ) (V – b) = RT
2
a ab
or PV = RT + Pb – __ + ___2
V V
a
= RT + Pb – __
V ( ab
___
V2
is negligible
)
Substituting the above value of PV in Eqn. (3)
3.20 Heat and Thermodynamics

a a a a
( ) ( )
W = RT2 + P2b – ___ – RT1 + P1b – ___ + ___ – ___
V2 V1 V1 V2
2a 2a
= R(T2 – T1) – b(P1 – P2) + ___ – ___
V1 V2
Now, if from the ideal gas equation PV = RT, the approximate values of V1 and
RT1 RT2
V2 are assumed to be V1 = ____ and V2 = ____ respectively, then
P1 P2

2aP1 2aP2
W = R(T2 – T1) – b(P1 – P2) + _____ – _____
RT1 RT2

Now, since T1 and T2 are nearly equal, hence assuming T1 – T2 = dT and


T1 ª T2 = T (initial temperature), we get
2a
W = – RdT – b(P1 – P2) + ___ (P1 – P2)
RT

[
2a
]
= (P1 – P2) ___ – b – RdT
RT
(4)

Since the porous plug is insulated hence neither the gas absorbs heat nor it
rejects heat. As a result, the total work done by the gas is obtained from the kinetic
energy of its molecules due to which the temperature of gas decreases by dT. If
molar specific heat of the gas at constant volume is CV, the decrease in internal
kinetic energy of gas due to fall in temperature by dT is
dU = CV (T2 – T1) = CV dT (5)
Now, from the law of conservation of energy
W = dU

or (2a
)
(P1 – P2) ___ – b – RdT = CV dT
RT

or
2a
(
(CV + R) dT = (P1 – P2) ___ – b
RT )
Since CP – CV = R
or CV + R = CP

\
2a
(
CP dT = (P1 – P2) ___ – b
RT )
or
1 2a
( )
dT = ___ ___ – b (P1 – P2)
CP RT
(6)
Real Gases, Liquefaction of Gases and Production and... 3.21

The above equation gives the expression for cooling of a real gas in Joule
Thompson effect. In this expression, since P2 < P1 therefore P1 – P2 is always
positive. Obviously, in Joule-Thompson effect, the change in temperature of a real
gas is directly proportional to the pressure difference P1 – P2 across the porous
plug.
Now, from Eqn. (6), it is clear that
2a
RT
2a
Rb
2a
( )
(i) If ___ > b or T < ___, the quantity ___ – b is positive, hence dT will be
RT
positive i.e., the temperature decreases with decrease in pressure. In other
words, the gas on passing through the porous plug gets cooled.
2a
RT
2a
Rb
2a
( )
(ii) If ___ < b or T > ___, the quantity ___ – b is negative, hence dT will be
RT
negtive i.e., the temperature increases with decrease in pressure. In other
words, the gas after passing through the porous plug gets heated.
2a
RT
2a
Rb (2a
)
(iii) If ___ = b or T = ___, the quantity ___ – b = 0, hence dT = 0, i.e., there is
RT
no change in temperature of gas on passing through the porous plug. This
temperature is called the temperature of inversion
2a
Ti = ___ (7)
Rb
The temperature of inversion of hydrogen and helium are below the ordinary
temperature. Therefore, these gases (H2 and He) show heating effect in adiabatic
throttling process. If these gases (H2 and He) are first cooled below their
temperature of inversion and then passed through the porous plug, they also show
the cooling effect.

3.13 TEMPERATURE OF INVERSION


The temperature of inversion of a gas is the temperature below which a gas gets
cooled in Joule-Thomson expansion and above which the gas gets heated, i.e., at
this temperature, the sign of Joule-Thomson effect is reversed. In other words, the
gas must be initially cooled below its temperature of inversion to produce cooling
by Joule-Thomson effect.
The expression for temperature of inversion given by Eqn. (7) is an approximate
expression. The actual expression is obtained by equating the quantity
a
(
[T(∂V/∂T)P – V] to zero.* From van der Waal’s gas equation P + ___2 (V – b)
V )
= RT, at T = Ti i.e., 2aV (V – b)2 – RTV 3b = 0, at T = Ti.

*
Refer to Section 3.14.
3.22 Heat and Thermodynamics

or 2aV (V – b)2 = RTiV 3b

2a (V – b)2
or Ti = _________ (8)
RV 2b
Robeck and Oesterberg experimentally
observed the Joule-Thomson effect at
400
different initial pressures. They found that
there are two temperatures of inversion 300 225°C

Temperature (in °C)


of a gas for a given initial pressure and
200
the values are different at different initial
pressures. A graph plotted between the 100
Cooling Heating
initial pressure and the temperature of 0
inversion is a parabola (Fig. 3.14). It is
called the inversion curve. At a temperature – 100 –120°C
and pressure corresponding to the left – 200
of this curve, cooling is produced while 0 100 200 300 400
corresponding to the right of this curve, Pressure (in atmosphere)
heating is produced. Figure 3.14 shows Fig. 3.14 Temperature of inversion
the inversion curve for nitrogen. It is clear curve for N2
that if the initial pressure of nitrogen is
200 atmosphere, then cooling is produced in Joule-Thomson expansion when the
initial temperature is in the range from – 120°C to 225°C. But if initial temperature
is below – 120°C or above 225°C, heating is produced. Thus, the temperatures of
inversion of nitrogen at 200 atmospheric pressure are – 120°C and 225°C.

3.14 PRINCIPLE OF JOULE THOMSON’S POROUS


PLUG EXPERIMENT
Joule’s experiment was not capable of measuring precisely the extremely small
temperature changes in a free expansion of a gas. However, Joule and Thomson
(later Joule and Kelvin) devised another better experiment during 1850-1860.
They devised their experiment in such a way that the temperature change due to
expansion of a gas would not be masked by the comparatively large heat capacity
of the surroundings. The result of this experiment provided information about
intermolecular forces. Further, temperature drop produced in the experiment
could be used in the liquefaction of gases such as hydrogen and helium.
The schematic representation of the experimental set up is shown in Fig. 3.15.
A cylindrical tube, insulated to prevent any transfer of heat to the surroundings
is fitted with two pistons and a porous plug which is capable of allowing gas to
flow slowly through it. The tube A is initially filled with a certain amount of gas
at temperature T1, volume V1 and pressure P1; tube B is empty. The gas is then
Real Gases, Liquefaction of Gases and Production and... 3.23

P1 P2

A B
V1 V2

T1 T2

Porous Plug

Fig. 3.15 Schematic arrangement of porous plug experiment

allowed to flow slowly through the plug in such a way that its pressure in A is kept
constant at P1 by the movement of the piston towards the plug. At the same time
piston B is adjusted in such a way that the low pressure P2(< P1) is kept constant.
Let the final volume in B, after all the gas has streamed through the porous plug,
be V2 and its temperature T2. The significant datum obtained in this experiment
is the change in temperature due to flow of the gas through the porous plug (by
measuring temperatures T1 and T2).
The whole system is insulated, so there is not heat transfer to the surroundings
and therefore Q = 0. Therefore, the change in internal energy is equal to the work
done by the system. The total work done is
W = P1V1 – P2V2 (1)
Then DU = U2 – U1 = P1V1 – P2V2
or (U2 + P2V2) – (U1 + P1V1) = 0
or H2 – H1 = 0
or DH = 0 (2)
This shows that the Joule-Thomson experiment is carried out under constant
enthalpy conditions.
When the gas involved is perfect, H is a function of T only and therefore DH = 0
implies that DT = 0 or T2 = T1 (no temperature change for a perfect gas). For an
imperfect gas it generally depends on whether T2 > T1 or T2 < T1> If T2 < T1, the
gas will be cooled and if T2 > T1, the gas will be heated. The crucial temperature
is called the Joule-Thomson inversion temperature. Above this temperature, there
will be heating and below this temperature there will be cooling upon Joule-
Thomson expansion.
For most gases, the Joule-Thomson inversion temperature lies above the room
temperature so, no precooling is necessary before they are allowed to expand
3.24 Heat and Thermodynamics

through a porous plug. Helium and hydrogen are exemptions; these are heated
when Joule-Thomson expansion occurs at room temperature (They require
precooling for liquefaction).
The Joule-Thomson coefficient m is defined as the change in temperature per
unit change in pressure when the enthalpy is constant i.e.,

( )
∂T
m = ___
∂P H
(3)

From this equation it follows that if the gas cools in the process, the Joule-
Thomson coefficient m is positive, because the pressure always decreases in the
experiment. Conversely, a negative Joule-Thomson coefficient implies an increase
in temperature. At a definite temperature for each gas, called the inversion point,
the Joule-Thomson effect must therefore change sign.
If we consider now H as a function of temperature and pressure, the total
differential of H is

( )
∂H
dH = ___
∂T P
∂H
dT + ___
∂P ( ) T
dP (4)

∂H
∂T( )
But ___
P
= CP, therefore

∂H
dH = CP dT + ___
∂P ( ) T
dP (5)

Since in Joule-Thomson experiment, there is no change in enthalpy, i.e.,


dH = 0, so
∂H
∂P( )
0 = CP (dT)H + ___
T
(dP)H

or
∂T
0 = CP ___
∂P ( ) ( ) H
∂H
+ ___
∂P T

or ( ) ∂H
___
∂P T
∂T
= – CP ___
∂P ( ) H
= – CP m (6)

Since dH = 0, it follows from Eqn. (4) that

( ) ( )( )
∂T
___
∂P H
∂H
= – ___
∂P T
∂H
___
∂T
–1

P
(7)

It can be shown* that

( ) ∂H
___
∂P T
∂V
= V – T ___
∂T ( ) P
(8)

*
H = U + PV fi dH = dU + PdV + VdP = TdS + VdP
∂H
fi ___
∂P ( ) T
( )
∂S
= V + T ___
∂P T
( )
∂V
= V – T ___
∂T P
∂S
using Maxwell’s relation ___
∂P ( ) ( )
T
∂V
= – ___
∂T P
Real Gases, Liquefaction of Gases and Production and... 3.25

and we already know that (∂H/∂T)P = CP, therefore Eqn. (7) gives

T(∂V/∂T)P – V
( )
∂T
___
∂P H
= ____________
CP

=m (Joule-Thomson coefficient) (9i)

If CP is assumed to be constant over a small temperature range, then Eqn. (9i)


can be written as
T(∂V/∂T)P – V
DT = ____________ DP (9ii)
CP
This is the equation for the differential Joule-Thomson effect, and is in
agreement with experiments.
At the Joule-Thomson inversion temperature, m = 0. Below this temperature m is
negative and above, it is positive. Most gases at room temperature have a positive
Joule-Thomson coefficient. They do not require precooling for liquefaction by
Joule-Thomson expansion, since (∂T/∂P)H > 0 fi gas temperature falls with
pressure. On the other hand, (∂T/∂P)H < 0 fi when the gas goes from a higher to
a lower pressure in a Joule-Thomson process, its temperature rises.

Case (i) Perfect Gas


The equation of state is PV = RT

fi ( )∂V
___
∂T P
R
= __
P

or ( )
∂V
T ___
∂T P
RT
= ___ = V
P

or
∂V
T ___
∂T ( ) P
–V=0

\ m=0 (From Eqn. (9i).


i.e., Joule-Thomson coefficient for a perfect gas is zero. Also DT = 0 from
Eqn. (9ii).

Case (ii) Real Gas

( P + ___Va ) (V – b) = RT
2
(10)

Differentiating it with respect to T taking P constant, we have

( P + ___Va ) ( ___∂V∂T ) – ___2aV ( ___∂V∂T ) (V – b) = R


2 P 3 P
(11)
3.26 Heat and Thermodynamics

or ( )
∂V
___
∂T P a 2a
R
= ________________
P + ___2 – ___3 (V – b)
V V
R
= _______________
RT 2a
_____ – ___ (V – b)
V – b V3
R(V – b)
= ______________
2a
RT – ___3 (V – b)2
V
RT (V – b)
( )
∂V
T ___
∂T P
= ______________
2a
RT – ___3 (V – b)2
V
RT (V – b)
= _________ ( b << V)
2a
RT – ___
V
V–b 2a
[
= _______ = (V – b) 1 – ____
1–
2a
____ RTV ] –1
(12)
VRT
Expanding binomially and neglecting higher order terms, we have

( )
∂V
T ___
∂T P
[ 2a
= (V – b) 1 + ____
VRT ]
2a
= V – b + ___
RT ( neglecting ____
VRT )
2ab

or ( )
∂V
T ___
∂T P
2a
– V = ___ – b
RT
(13)

fi (
1 2a
m = ___ ___ – b
CP RT ) (14)

2a 2a
Thus, if ___ > b or T < ___ then m (= (∂T/∂P)H) is positive. But since ∂P is
RT Rb
negative (pressure on emergent side of the porous plug is lower) so, ∂T will be
negative and the gas will cool on passing through the porous plug.
If 2a/RT < b or T > 2a/Rb, the gas will be heated up. At T = 2a/Rb, there will
be no change in temperature. Therefore, T = 2a/Rb = Ti (inversion temperature).
For most of the gases, the ordinary working temperatures are below Ti ,
hence they show a cooling effect. Ti for H2 and He are much below the ordinary
temperatures, hence, they show a heating effect. If, however, these gases are
Real Gases, Liquefaction of Gases and Production and... 3.27

precooled below Ti before undergoing Joule-Thomson expansion, they will also


show a cooling effect. It may be mentioned here that the Boyle temperature at
a
which the gas obeys Boyle’s law is TB = ___. Therefore, Ti = 2TB i.e., temperature
Rb
of inversion = 2 × Boyle’s temperature.

3.15 CHANGE IN TEMPERATURE IN AN ADIABATIC CHANGE


From Maxwell’s thermodynamic relations

( ) ( )
∂T
___
∂V S
∂P
= – ___
∂S V
(see Chapter 5).

or ( ) ( )
∂T
___
∂V S
T ∂P
= – __ ___
T ∂S V

( )
∂P
= – T ___
∂Q V
(1)

( T∂S = ∂Q)

since the pressure of the system increases at a constant volume hence the
∂P
( )
∂Q V
∂T
( )
quantity ___ is positive and the quantity ___ is negative. It means that at
∂V S
a constant entropy (i.e., in an adiabatic process), the temperature decreases on
increasing volume, i.e., cooling is produced in adiabatic expansion.
Since (∂Q)V = CV dT
\ From Eqn. (1)

( )
∂T
___
∂V S
T ∂P
( )
= – ___ ___
CV ∂T V
(2)

Now, if pressure coefficient of gas is b, then


increase in pressure
b = _________________________________
initial pressure × increase in temperature

( )
1 ∂P
= __ ___
P ∂T V

or ( )
∂P
___
∂T V
= Pb

Hence, from Eqn. (2)

( )
TPb
dT = – ____ dV
CV
(3)
3.28 Heat and Thermodynamics

It is thus clear that in adiabatic expansion, temperature decreases with increase


in volume and the above Eqn. (3) represents the decrease in temperature in
adiabatic expansion.
Similarly, from Maxwell’s thermodynamic relation

( ) ( )
∂T
___
∂P S
∂V
= ___
∂S P

( ) T ∂V
= __ ___
T ∂S P

or ( ) ( )
∂T
___
∂P S
∂V
= T ___
∂Q P
(4)

Since, on imparting heat to the system at a constant pressure, its volume


∂V
∂Q P( ) ∂T
( )
increases i.e., the quantity ___ is positive, hence the quantity ___ will also
∂P S
be positive which means that at a constant entropy, (i.e., in adiabatic process),
temperature increase with the increase in pressure or temperature decreases with
decrease in pressure i.e., heating is produces in adiabatic compression and cooling
is produced in adiabatic expansion.

Since (∂Q)P = CP dT
\ From Eqn. (4)

( )
∂T
___
∂P S
T ∂V
( )
= ___ ___
CP ∂T P
(5)

Now, if the coefficient of volume expansion of gas is a, then


increase in volume
a = _________________________________
initial volume × increase in temperature
1 ∂V
( )
= __ ___
V ∂T P
or ( )
∂V
___
∂T P
= Va

Now, from Eqn. (5)

( )
TVa
dT = ____ dP
CP
(6)
It is clear that in adiabatic expansion, temperature decreases with the decrease
in pressure and the above Eqn. (6) represents the decrease in temperature in
adiabatic expansion.
The Joule-Thomson (also known as Joule-Kelvin) effect discussed in the
previous section, is an adiabatic throttling process in which cooling or heating
.

Real Gases, Liquefaction of Gases and Production and... 3.29

is produced depending on whether the initial temperature of the gas is below or


above the temperature of inversion.

3.16 DISTINCTION BETWEEN JOULE’S EXPANSION, JOULE‐


THOMSON’S EXPANSION AND ADIABATIC EXPANSION
In all the three Joule’s expansion, Joule-Thomson’s expansion and adiabatic
expansion, the system is thermally isolated from its surroundings. Consequently,
in all the three processes, there is no exchange of heat with the surroundings.
Even then, these processes differ in the following respects:
(i) In Joule’s expansion, the system is mechanically isolated from the
surroundings, hence the total external work done on the system or by the
system is zero (i.e., W = 0). In Joule-Thomson’s expansion, the system is
not mechanically isolated from its surroundings and pressure of the gas in
it is kept constant before and after the expansion as P1 and P2 respectively,
hence in this expansion external work is done by the system and on the
system (net external work done by the system is W = P2V2 – P1V1). In
adiabatic expansion, the system is neither isolated from its surroundings,
nor its pressure is kept constant, hence in this expression, only the external
work is done against the atmospheric pressure by the system (W = Ú PdV).
(ii) In Joule’s expansion, there is no change in internal energy (dU = 0). In
Joule-Thomson expansion, internal energy decreases (dU = P1V1 – P2V2).
In adiabatic expansion also, the internal energy decreases (dU = – Ú PdV).
(iii) In Joule’s expansion, there is no change in temperature of an ideal gas,
whereas there is always cooling in a real gas because the internal kinetic
energy decreases because of internal work done by the gas. In Joule-
Thomson’s expansion, although there is no change in temperature of an
ideal gas, but a real gas shows cooling when P2V2 > P1V1 or heating when
P2V2 < P1V1. In adiabatic expansion, the external work is obtained from its
own internal energy due to which there is always a decrease in temperature
of the system.

3.17 ADIABATIC DEMAGNETISATION


The production of low temperatures i.e., temperatures far below the freezing
point of ice was a subject of great scientific interest. The quest for reaching
temperatures close to the absolute zero (0°K) was the chief interest of the
physicists. The adiabatic expansion of a real gas was found to give rise to cooling
of a gas. When a gas passes through an orifice, or a set of orifices from a region
of constant high pressure to a region of constant low pressure adiabatically, it is
said to undergo throttling or Joule-Thomson expansion. A number of gases were
3.30 Heat and Thermodynamics

cooled to their boiling points. Of course, the gases, initially, should be at their
inversion temperatures. The gases which could be cooled to their boiling points
were hydrogen (33 K), oxygen 82K), Nitrogen (77 K) and finally helium (4.2 K).
Helium remains as liquid even below 1 K. The production of temperatures below
1K is not possible by Joule-Thomson process. Adiabatic demagnetization is found
to be highly useful in the production of temperatures below 1 K.

3.17.1 Principle
A paramagnetic substance (or salt) can be treated as a thermodynamic system
containing atoms which behave like small magnets, all lying in a disorderly
manner. When it is magnetized, these groups of dipoles set themselves parallel to
the lines of force, an external work is done on the system. If this already magnetized
substance is suddenly demagnetized, the axes of atomic magnets (due to thermal
agitation) tend to resume their natural disorderly state. Now work will be done by
the substance drawing energy from the substance itself and consequently cooling
will be produced. This is known as magnetocaloric effect.
The entropy S of a paramagnetic system is a function of B/T and is given by
m 2 mB
dS
___
dB ( )
= – Nk ___ B sec h2 ___
kT ( )
kT
(1)

Since sec h2 (mB/kT) is always positive, the entropy S decreases with increasing
B for a given T.
Figure 3.16 shows entropy S of a paramagnetic system for two fields B1 and
B2 (B1 < B2) as a function of temperature. We first increase the magnetic field

B1
a

B2
S

c b B1< B2

Tf Ti T

Fig. 3.16 Principle of adiabatic demagnetization


Real Gases, Liquefaction of Gases and Production and... 3.31

from B1 to B2, under isothermal condition, at a initial temperature Ti (from state


a to state b). During this process, there is a heat transfer from the paramagnetic
specimen to the heat bath. The specimen is then thermally isolated and the
magnetic field is reduced very slowly (adiabatic demagnetization) under adiabatic
conditions from B2 to B1. In this process, the entropy S of the specimen remains
constant (This is illustrated in the change from state b to state c), leading to the
lower temperature Tf (cooling by adiabatic demagnetization).

3.17.2 Theory
When a paramagnetic salt is placed in a magnetic field of strength H, the intensity
of magnetization per gm-mole I of the system change by dI. The work done by
the field on the substance = HdI.
or, work done by the substance = – HdI.
From first law, heat supplied
dQ = dU + dW (2)
But here
dW = PdV – HdI \ dQ = dV + PdV – HdI (3)
Since change in volume of a solid for small changes in temperature is negligible
i.e., dV = 0, thus
dQ = dU – HdI (4)
From second law
dQ = TdS (5)
where dS is change in entropy
\ TdS = dU – HdI
or dU = TdS – HdI (6)
Now, entropy S, intensity of magnetization I and internal energy U are functions
of absolute temperature T and magnetic field strength H. Consequently, we can
write

( )
∂S
dS = ___
∂T H
( )
∂S
dT + ___
∂H T
dH

( )
∂I
dI = ___
∂T H
( )
∂I
dT + ___
∂H H
dH (7)

( )
∂U
dU = ___
∂T H
( )
∂U
dT + ___
∂H H
dH
3.32 Heat and Thermodynamics

Substituting the values of dS, dI and dU in Eqn. (6), we get

( )
∂U
___
∂T H
( )
∂U
dT + ___
∂H H
dH = T {( ) ( ) }
∂S
___
∂T H
∂S
dT + ___
∂H T
dH

{ ( ) ( ) }
∂I
+ H ___
∂T H
∂I
dT + ___
∂H H
dH (8)

Comparing coefficients of dT on both sides,

( ) ( ) ( )
∂U
___
∂T H
∂S
= T ___
∂T H
∂I
+ H ___
∂T H
(9)

Comparing coefficients of dH on both sides of Eqn. (8)

( ) ( ) ( )
∂U
___
∂T H
∂S
= T ___
∂H T
∂I
+ H ___
∂H T
(10)

As U is a function of state of the system, dU must be a perfect differential


i.e.,

[ ( )] [ ( )] ∂ ___
___ ∂U
∂H ∂T H T
∂ ∂U
= ___ ___
∂T ∂H T H

or [ { ( ) ( ) }] [ { ( ) ( ) }]

___
∂H
∂S
T ___
∂T H
∂I
+ H ___
∂T H T

∂T
∂S
= ___ T ___
∂H T
∂I
+ H ___
∂H T H

This yields

( ) ( )
∂I
___
∂T H
∂S
= ___
∂H T
(11)

Substituting this in Eqn. (10) gives

( ) ( ) ( )
∂U
___
∂H T
∂I
= T ___
∂T H
∂I
+ H ___
∂H T
(12)

Using the above equation, we can calculate the internal energy (or the entropy)
in an arbitrary field from the internal energy (or entropy) in zero field at the same
temperature when the magnetization curves at different temperatures are known.
From Eqns. (4) and (7), we have

dQ = {( )
∂U
___
∂T H
∂U
∂H( ) } {( ) ( ) }
dT + ___
T
dH – H
∂I
___
∂T H
∂I
dT + ___
∂H T
dH

= [( )
∂U
___
∂T H
∂I
∂T{ ( ) ( )} ]
dT + T ___
H
∂I
+ H ___
∂H T
dH

{( ) ( ) }
–H
∂I
___
∂T H
∂I
dT + ___
∂H T
dH (using Eqn. 12)
Real Gases, Liquefaction of Gases and Production and... 3.33

\ dQ = {( ) ( ) }
∂U
___
∂T H
∂I
– H ___
∂T H
( )
∂I
dT + T ___
∂T H
dH (13)

fi Thermal capacity at constant field


∂Q
∂T( ) ( ) ( )
CH = ___
T
∂U
= ___
∂T H
∂I
– H ___
∂T H
(14)

or
∂I
dQ = CH dT + T ___
∂T ( ) H
dH (15)

During adiabatic demagnetization, no heat enters or leaves the system (dQ = 0)


so Eqn. (15) gives
∂I
CH dT + T ___
∂T ( ) H
dH = 0 (16)

Now, according to Curie’s law, molar susceptibility


C
c = __ (17)
T
C is Curie’s constant.
M IV
But c = __ = ___
H H
where V is the volume occupied by 1 gm mole of the substance and M is molar
magnetization. Therefore,
Hc C H
I = ___ = __ . __ (using Eqn. 17)
V T V
KH
= ___ (18)
T
C
where K = __ is the Curie constant per unit volume.
V
Differentiating Eqn. (18) w.r.t. T at constant field H,

( )
∂I
___
∂T H
K
dH = – __2 H
T
(19)

Substituting it in Eqn. (16), we get


K
( )
CH dT + T – __2 HdH = 0
T
KH
or TdT = ___ dH
CH
Integrating
Tf 0
KH
Ú TdT = Ú ___
C
dH (20)
Ti H H
3.34 Heat and Thermodynamics

or [ ]
T__2
2
Tf

Ti
K H2
[ ]
= ___ ___
CH 2
0

K
or Tf2 – Ti2 = – ___ H2
CH

or
[ K H
Tf = Ti 1 – ___ __
CH Ti ( )]
2 1/2
(21)

It may be noted that Tf < Ti i.e., the temperature of the salt falls during adiabatic
demagnetization. The fall in temperature is greater for larger values of the
magnetizing field H and low values of initial temperature Ti.

3.17.3 Experimental Method


Figure 3.17 gives the apparatus used in
the experiment. The paramagnetic salt
(gadolinium sulphate) is suspended in
vessel surrounded by liquid helium cooled
to ~1°K contained in a Dewar flask D1
which in turn is surrounded by a Dewar
flask D2 containing liquid H2.
The whole arrangement is placed
between the poles of a strong electromagnet,
providing a magnetic field of the order of
10 K-gauss.
The following procedure is now
adopted: Fig. 3.17
(i) The magnetic field is switched on
so that the specimen is magnetized.
(ii) The heat produced during magnetization is conducted away by the gaseous
helium in A to the liquid He in D1. Thus, the specimen in A is left both cold
and highly magnetized. It corresponds to isothermal magnetization.
(iii) The He-gas from the cylinder A is now pumped off with a high vacuum
pump. The specimen is now thermally isolated from D1 and D2.
(iv) The magnetic field is now switched off. Thus, instantaneous adiabatic
demagnetization of the salt takes place and its temperature falls to
~ 0.25 K.
The temperature of the specimen is determined by fitting a coaxial solenoid
coil round the tube A and measuring the self inductance and hence susceptibility
of the salt with the help of an a.c. bridge at the beginning and the end of the
experiment. If c1 is the susceptibility at temperature T1 of the He-bath and c2, the
Real Gases, Liquefaction of Gases and Production and... 3.35

susceptibility after adiabatic demagnetization at temperature T2,then according to


Curie’s law
c1 T2
__ = __
c2 T1
c1
or T2 = __ T1
c2

whence T2 can be calculated, which is called the Curie temperature or magnetic


temperature. From it, the Kelvin temperature can be determined.

3.18 LIQUEFACTION OF GASES AND APPROACH


TO ABSOLUTE ZERO
The science of the production and use of low temperatures is known as cryogenics.
The methods employed for the liquefaction of gases by cooling can be classified
as follows:
1. The rapid evaporation of volatile liquids.
2. The adiabatic expansion of cold compressed gases.
3. The Joule-Thomson effect.
4. The adiabatic demagnetization.
We briefly describe these methods below.

3.18.1 The Rapid Evaporation of Volatile Liquids


When a liquid evaporates, latest heat of vaporization has to be supplied to it. If
the liquid is thermally isolated, this heat comes from the liquid itself. As a result
the liquid is cooled.
We know that a gas can be liquefied only if it has been cooled below its critical
temperature Tc, Table 3.1. By rapidly evaporating one liquefied gas it is possible
to obtain a temperature low enough to liquefy another gas. This is called the
cascade process. It was first used by Cailletet and by Pictet in 1877 to liquefy
oxygen (Tc = – 119 C). Kamerlingh-Onnes, at Leyden, set up a cascade system
using methyl chloride (Tc = 143°C > room temperature, liquefied by compression,
cooled to – 90°C by evaporation at low pressure) and ethylene (Tc = 10°C, cooled
to – 140°C, by evaporation at low pressure) and was able to produce oxygen in
quantity. His apparatus is shown in Fig. 3.18.

3.18.2 The Adiabatic Expansion of Cold Compressed Gases


In all gases an approximately reversible adiabatic expansion (dS = 0) of a gas
against a piston produces cooling. We can write
3.36 Heat and Thermodynamics

Fig. 3.18 Cascade process

( ) ( )( )
∂T
___
∂V S
∂T
= – ___
∂S V
∂S
___
∂V T

( )
T ∂p
= – ___ ___
CV ∂T V

where all the terms on the right are positive. Therefore, if V increases, T
decreases.
The physical reason for cooling is simply that in adiabatic expansion the work
is done on the moving piston at the expense of the kinetic energy of the molecules,
resulting in temperature decrease.
The amount of cooling can be easily calculated. For the adiabatic process
Tf = Ti (pi/pf)(1 – g)/g
If pi = 51 atm, pf = 1 atm, Ti = 27°C and g = 1.4, we have
Tf = 300 × (51)– 0.4 / 1.4 = 97.4°K
Ti – Tf = 300 – 97.4 = 202.6°K = 202.6°C
The apparatus of Claude is shown in Fig. 3.19. Purified air is compressed to 40
atm and passed through the tube A into the piston-cylinder arrangement B where
it expands doing external work. The cooled and expanded air moves through the
liquefier and finally passes along the outer tube of the heat exchanger A, cooling
the incoming gas. This goes on until liquid air is formed.
Real Gases, Liquefaction of Gases and Production and... 3.37

A Adiabatic
expansion
B
Piston
Heat
exchanger

Liquefier

Fig. 3.19 Claude’s apparatus

3.18.3 The Joule-Thomson Effect


For cooling to result in the Joule-Thomson expansion we must have m > 0 or
T < Ti. The amount of cooling produced is given by (using Eqns. 9(i) and (14) of
3.14).
pi – pf
Ti – Tf = _____
cp T [( ) ]
∂v
___
∂T p
–v
pi – pf 2a
= _____
cp [ ]
___ – b
RTi
For example, for oxygen using pi – pf = 150 atm @ 150 × 106 dynes per cm2
cp = 7 cal/mole. °K, Ti = 27°C = 300°K, a = 1.32 litre2 atm per mole2, and
b = 3.12 × 10– 2 litre per mole, we have
2a
___ 2 × 1.32 × 1012
– b = ________________ – 31.2
RTi 2 × 4.2 × 107 × 300
= 73.6 cm3 per mole
150 × 106 × 73.6
Ti – Tf = ______________ = 37.5°K
7 × 4.2 × 107
= 37.5°C.
If Ti is taken to be 75°K instead of 300°K, we get
Ti – Tf = 198°K = 198°C.

(a) Hampson’s Regenerative Cooling Method and Linde’s Method


Joule-Thomson cooling is small at room temperatures but increases as the initial
temperature is lowered. This fact has been used in the regenerative cooling
3.38 Heat and Thermodynamics

method given by Hampson of England. The principle is shown in Fig. 3.20.


The compressed air after passing through a water bath, which removes the
compression heat, passes to the heat exchanger. Air at a pressure of 150 atm is
allowed to undergo Joule-Thomson expansion through a porous plug or throttle
valve. The expanded air returns to the compressor through the outer jacket of
the heat exchanger (counter-current heat exchanger), cooling the incoming air
(regenerative cooling). As the temperature of the incoming air is lowered, the
Joule-Thomson cooling increases rapidly. After a few cycles the air gets liquefied
under Joule-Thomson effect alone.
Water bath

20°C
150 atm

Heat
Compressor exchange

Porousplug
or
throttle
valve

1 atm – 180°C

Fig. 3.20 Regenerative cooling method of Hampson [After: B.K. Agarwal, Thermal Physics,
Lokbharti Publications, Allahabad (India) (1988)]

Linde independently developed a similar method. The difference is that the


Joule-Thomson expansion takes place from 150 atm to 40 atm and the gas needs
to be compressed back from 40 atm to 150 atm, instead of 150 atm. If we assume
isothermal compression the work involved is proportional to ln (pf /pi). Since
cooling is proportional to (pi – pf), we have
ln (pf /pi)LINDE ln (150/40)
_____________ = __________
ln (pf /pi)HAMPSON ln (150/1)
= 0.26
Real Gases, Liquefaction of Gases and Production and... 3.39

(Ti – Tf)LINDE 150 – 40


____________ = ________
(Ti – Tf)HAMPSON 150 – 1

= 0.74

Thus, the Linde process gives about


3/4 the cooling for 1/4 the work, that is,
it is more efficient.
Figure 3.21 gives a flow diagram
of the Linde process. The pump P1
compresses the air from 1 atm to 40 atm.
The circulating pump P2 compresses
the air from 40 atm to 150 atm. The
compressed air gets cooled in ammonia
bath E1 passes through the heat
exchanger E2, and expands through the
throttle value V1, becoming liquid at 40
atm and – 183°C. The unliquefied gas
rises up through E2 back to P2, cooling
the incoming air on the way.

(b) Claude’s Process and


Kamerlingh-Onnes Method
Fig. 3.21 Linde process
A combination of methods 2 and 3 is
more successful. The adiabatic
reversible expansion is used to
achieve a temperature below Ti
and then the Joule-Thomson effect
completes the liquefaction. Claude
process shown in Fig. 3.22 is based
on this principle.
Air at 40 atm is cooled in
exchanger E1 and divided into
two streams: 80% goes into the
adiabatic expansion cylinder-
piston arrangement, gets cooled
and moves up the heat exchanger
E2 to cool the other 20%. This
cooled 20% stream undergoes
Joule-Thomson expansion at the
Fig. 3.22 Claude process
3.40 Heat and Thermodynamics

throttle valve V, a part of it gets liquefied and the remainder rises up the exchanger
E2 to cool the incoming air.
It is useful to know the points of distinction between adiabatic and Joule-
Thomson expansion, Table 3.2.

Table 3.2 Points of distinction between adiabatic and Joule-Thomson expansions.


Adiabatic expansion Joule-Thomson expansion
1. Net external work done by the gas. 1. No net external work done by the gas.
2. Entropy remains constant. 2. Enthalpy remains constant.
3. Cooling produced for perfect gas. 3. Cooling not produced for perfect gas.
4. Cooling produced for real gas at all 4. For real gas heating above Ti, no effect
initial temperatures. at Ti, and cooling below Ti.
5. Cooling decreases as initial temperature 5. Cooling increases as initial temperature
decreases. decreases.
6. Reversible process. 6. Irreversible process.

Liquefaction of hydrogen is difficult because its critical temperature is 33°K,


which is much below the liquid air temperature 78°K. Its inversion temperature
is 194°K. So, it can be precooled to below 194°K with the help of liquid air and
then liquefied by Joule-Thomson effect. Kamerlingh-Onnes designed a simple
apparatus on this principle, Fig. 3.23. Purified hydrogen is passed through a bath

Fig. 3.23 Kamerlingh-Onnes method for hydrogen liquefaction by Joule-Thomson effect


Real Gases, Liquefaction of Gases and Production and... 3.41

of liquid air boiling under reduced pressure at – 208°C. The hydrogen at 150 atm
and – 208°C undergoes Joule-Thomson expansion to a pressure of 1 atm and gets
liquefied.
The inversion temperature for helium is about – 240°C or 33°K. Kamerlingh-
Onnes liquefied it in the same way by precooling to – 258°C by the aid of a bath
of hydrogen boiling under reduced pressure.

3.18.4 The Adiabatic Demagnetization


This method was first perfected by Giauque in 1933 to produce extremely low
temperatures, within a very small fraction of a degree of absolute zero.
Figure 3.24 shows the variation of entropy with temperature for a paramagnetic
salt for two values of external magnetic field, 0 and Hi. For H = 0, the sudden fall
in entropy at the Curie temperature TC corresponds to the onset of spontaneous
ordering.
A rare earth paramagnetic salt is first magnetized by applying a field Hi at an
initial temperature Ti (~1°K). By evaporating liquid He4 under reduced pressure,
a temperature of about 1°K can be
easily obtained. The heat evolved
during magnetization is absorbed by
the helium bath. The entropy falls
and the salt goes from state a to state i
(Fig. 3.24). Now the salt is thermally
isolated and demagnetized. For
slow demagnetization the process is
reversible, entropy remains constant,
and the temperature falls. The salt goes
from state i to state f with temperature
Tf. The lowest temperature attainable Fig. 3.24 The entropy vs. temperature graph for a
by demagnetization is effectively the paramagnetic salt at H = 0 and H = Hi
Curie temperature.
The experimental arrangement for adiabatic demagnetization is given
in Fig. 3.25. The salt A is placed in the chamber R which is immersed in the
helium bath. For isothermal magnetization R is filled with helium gas to provide
thermal contact with the liquid helium bath. After magnetization in a field of
about 10,000 oersted, R is evacuated, thermally isolating the salt. The field is now
reduced to zero. There is an accompanying drop in the temperature of the salt.

3.19 PRODUCTION OF VERY LOW TEMPERATURES


For the liquefaction of gases, it is required to produce temperatures below 0°C.
Apart from this, production of low temperature is important because: (i) many
3.42 Heat and Thermodynamics

Fig. 3.25 Adiabatic demagnization

physical properties of a substance show a specific behaviour at low temperatures;


(ii) the phenomenon of superconductivity is only possible at low temperatures.
Following are the major methods of production of very low temperatures:
1. By adding salt in ice,
2. By vaporization of liquids at low pressure,
3. By adiabatic expansion of gases,
4. By Joule-Thomson effect,
5. By regenerative cooling,
6. By adiabatic demagnetization.
1. By adding salt in iceWhen a soluble salt (such as common salt, NaCl) is
mixed in ice, the ice takes some heat from it and starts melting. The latest
heat required for melting of ice is obtained from the mixture. Moreover,
the salt dissolves in water which is formed by melting of ice, for which
the heat of solution required, is also taken from the mixture. As a result,
the temperature of mixture falls. As the amount of iced water in mixture
increases, the temperature of mixture decreases and a stage is reached when
Real Gases, Liquefaction of Gases and Production and... 3.43

no more salt dissolves and the mixture attains a minimum temperature.


The minimum temperature of mixture is different for different salts. For
example, on adding ammonium chloride, it reaches – 55°C and on adding
potassium hydroxide, it reaches – 65°C.
2. By vaporization of liquids at low pressure When a liquid vaporizes in
changing from liquid state to the gaseous state, it take the required latent
heat of vaporization from the liquid and its surroundings due to which the
temperature of liquid falls. More the latent heat of vaporization of liquid,
more is the cooling produced. Moreover, if the pressure at the liquid surface
is decreased, the boiling point of liquid will decrease due to which its
vaporization will be rapid and the temperature of liquid will reduce more.
Thus, the amount of cooling produced depends on the nature of liquid and
its rate of vaporization. This principle is used in a refrigerator in which a
low temperature is produced by the evaporation of liquid ammonia, liquid
methyl chloride, sulphur-dioxide, freon etc. under the reduced pressure.
3. By adiabatic expansion of gases We have read that when a gas expands
adiabatically, the gas does work against the external pressure due to which
the internal energy of the gas decreases and cooling is produced. By this
method, we can reach to a low temperature at which the gas freezes to a
solid. For example, if the tap of a cylinder containing CO2 gas at a high
pressure is suddenly opened, we get the gas in form of white snow on a
piece of cloth placed at the jet of gas.
4. By Joule-Thomson expansion When a compressed gas at an initial temperature
below its temperature of inversion is passed through a narrow nozzle from
a constant high pressure to a constant low pressure, the temperature of gas
decreases. The fall in temperature of a gas is directly proportional to the
pressure difference across the nozzle. Lower the initial temperature of gas
below its temperature of inversion, higher is the cooling of gas produced.
This is called the Joule-Thomson expansion.
5. By regenerative cooling For liquefaction of gases, the gas which has suffered
Joule-Thomson expansion is used to cool the coming gas before its suffers
the Joule-Thomson expansion. As a result, this after suffering Joule-
Thomson expansion gets further cooled. This process is repeated again
and again, hence, the process is called the regenerative cooling and by this
method, a very low temperature can be attained.
6. By adiabatic demagnetization When a paramagnetic substance is magnetized
at a low temperature, the molecular magnets of the substance align
along the magnetizing field in which work done by the magnetic field is
stored in the substance in form of its internal energy and the temperature
of substance increases. Now the substance is first cooled to the initial
3.44 Heat and Thermodynamics

temperature and then it is adiabatically demagnetized due to which the


temperature of substance decreases to a great extent. This method is used
to attain a temperature very near to absolute zero.

3.20 MEASUREMENT OF VERY LOW TEMPERATURES


To measure the temperatures below 0°C, following thermometers are used:
(i) Liquid thermometer,
(ii) Gas thermometer,
(iii) Resistance thermometer,
(iv) Thermocouple thermometer,
(v) Vapour pressure thermometer,
(vi) Magnetic thermometer.
(i) Liquid thermometer The freezing point of mercury is – 39°C, hence, a
mercury thermometer can be used to measure a temperature upto – 39°C.
To measure a temperature below it, mercury can be replaced by toluene
(upto – 95°C), alcohol (upto – 112°C), pentane (upto – 132°C) and
isopentane (upto – 160°C).
(ii) Gas thermometer A constant volume hydrogen gas thermometer is used
to measure a temperature upto – 253°C and a constant volume helium gas
thermometer is used upto – 268.7°C.
(iii) Resistance thermometer A platinum resistance thermometer can measure a
temperature upto – 190°C to a great accuracy but the essential requirement
is that the platinum used must be in the pure state. To measure temperature
in the range of –190°C to – 258°C, lead resistance thermometer is used
instead of platinum resistance thermometer. In the temperature range
below it (upto – 272°C) an alloy resistance thermometer such as that of
constantan or phosphor bronze is used.
(iv) Thermocouple thermometer A copper-constantan or iron-constantan
thermocouple can measure a temperature upto – 200°C easily, but
to measure temperature below – 200°C (upto – 258°C), gold-silver
thermocouple is used.
(v) Vapour pressure thermometer The vapour pressure thermometers are much
sensitive for the measurement of low temperatures. The choice of vapour
used in a thermometer depends on the range of temperature to be measured
by it. Oxygen is used in the temperature range – 150°C to – 210°C, neon in
the temperature range – 246°C to – 249°C, hydrogen in temperature range
– 253°C to – 262°C and helium is used in the temperature range – 268°C
to – 272.29°C.
Real Gases, Liquefaction of Gases and Production and... 3.45

(vi) Magnetic thermometer This thermometer is used to measure a temperature


below – 272°C. Its principle is based on Curie law, according to which
the magnetic susceptibility of a paramagnetic substance is inversely
proportional to its absolute temperature.
Example 3.1 The volume of 1 g mole of gas at 0°C is 0.55 litre. Calculate its
pressure if the gas obeys (i) Ideal gas equation (ii) van der Waals equation. Given
a = 0.37 Nm4 mole–2, b = 43 cm3 mole–1 and R = 8.31 J mole–1 K–1.
Solution Given V = 0.55 litre = 0.55 × 10–3 m3
R = 8.31 J/(mole K).
T = 0°C = 273 K
a = 0.37 Nm4 mole–2,
b = 43 cm3/mole = 43 × 10–6 m3/mole
(i) From the ideal gas equation
PV = RT for 1 mole of gas
RT
or P = ___
V
8.31 × 273
\ P = __________
0.55 × 10–3
= 4.125 × 106 N/m2
(ii) From van der Waals equation

( P + ___Va ) (V – b) = RT for 1 mole of gas


2

RT a
\ P = _____ – ___2
V–b V
8.31 × 273 0.37
or P = _____________________ – ____________
(0.55 × 10 – 43 × 10 ) (0.55 × 10–3)2
–3 –6

8.31 × 273 ____________


0.37
= _________ –
507 × 10 –6
(0.55 × 10–3)2
= 4.475 × 106 – 1.223 × 106

= 3.252 × 106 N/m2


Example 3.2 The van der Waals constants of CO2 gas are a = 172 atm cm6 mole–2
and b = 0.002 cm3 mole–1. If R = 8.3 J mol–1 K–1, calculate its critical temperature,
Boyle temperature and temperature of inversion.
3.46 Heat and Thermodynamics

Solution Given a = 172 atm cm6 mole–2


= 172 × 10–7 Nm4 mole–2
b = 0.002 cm3 mole–1
= 2 × 10–9 cm3 mole–1
R = 8.3 Joule mol–1 K–1
Critical temperature
8a 8 × 172 × 10–7
TC = _____ = ________________
27Rb 27 × 8.3 × 2 × 10–9
= 307 K (= 34°C)
Boyle temperature
a 172 × 10–7
TB = ___ = ____________
Rb 8.3 × 2 × 10–9
= 1036.14 K (= 763.14°C)
Temperature of inversion
2a 2 × 172 × 10–7
Ti = ___ = ____________
Rb 8.3 × 2 × 10–9
= 2072.3K (= 1799.3°C)
Example 3.3 The van der Waals constants for a gas are a = 0.37 Nm4/mole2 and
b = 43 cm3/mole. Find its critical constants.
Solution Given a = 0.37 Nm4/mole2,
b = 43 × 10–6 m3/mole
\ Critical volume
VC = 3b = 3 × 43 × 10–6
= 1.29 × 10–4 m3/mole
Critical pressure
a 0.37
PC = ____2 = ______________
27b 27 × (43 × 10–6)2
= 7.4 × 106 N/m2
Critical temperature
8a 8 × 0.37
TC = _____ = ___________________
27Rb 27 × 8.31 × (43 × 10–6)
= 305.6 K or 32.6°C
Real Gases, Liquefaction of Gases and Production and... 3.47

QUESTIONS
1. Describe in brief the experiment performed by Andrews on CO2 and
discuss the results of the experiment.
2. Write the van der Waals equation of state. Compare the theoretical
isothermal curves obtained from this equation with the Andrews
experimental curves.
3. Write the van der Waals equation of state and derive expressions for critical
constants in terms of van der Waals constants a and b.
4. Obtain expressions for van der Waals constants a and b of a gas in terms
of critical constants.
5. What do you mean by Boyle temperature? Use van der Waals equation of
state to obtain an expression for Boyle-temperature of a gas.
6. Define critical temperature, Boyle temperature and temperature of
inversion of a gas and deduce relationship among them.
7. What is meant by temperature of inversion of a real gas? Obtain a expression
for it.
8. Show that Ti = 2TB and Ti = 6.75 TC where Ti, TB and TC are respectively
the temperature of inversion, Boyle’s temperature and critical temperature
of a gas.
9. Explain the principle of regenerative cooling of a gas.
10. What is cascade process of cooling? Explain the liquefaction of O2 by this
method with the help of a neat diagram.
11. What is Joule’s law for an ideal gas? Describe Joule’s experiment to
establish the existence of inter-molecular forces in a gas. What conclusions
were drawn from this experiment?
12. What is Joule-Thomson effect? Describe porous plug experiment and
discuss the results obtained from it.
13. What is Joule-Thomson effect? Obtain expression for Joule-Thomson
cooling produced in van der Waals gas and explain why at ordinary
temperatures hydrogen and helium gases show the heating effect while the
other gases show cooling effect.
14. Show that in adiabatic throttling process, the enthalpy of a system remains
constant. Obtain an expression for the change in temperature of a van der
Waals gas due to Joule-Thomson effect.
15. Describe the Linde’s method of liquefaction of air.
16. Describe Kammerlingh Onne’s method for liquefaction of hydrogen and
helium.
3.48 Heat and Thermodynamics

17. Describe with diagram the method of liquefaction of hydrogen and explain
the principle on which this method is based.
18. Write short notes on
(a) Boyle temperature
(b) Critical point
(c) Temperature of inversion
(d) Liquefaction of gases

OBJECTIVE TYPE QUESTIONS


1. A graph plotted for PV against P for an ideal gas is
(a) straight line parallel to P axis.
(b) a straight line parallel to PV axis.
(c) an inclined line passing through origin and making an angle of 45°
with P-axis.
(d) a hyperbola.
2. The isotherms of CO2 in Andrew’s experiment show that
(a) CO2 can be liquefied by increasing the pressure at any temperature.
(b) states are discontinuous.
(c) CO2 is a perfect gas.
(d) states are continuous.
3. The isothermals of a real gas are hyperbolic
(a) below the critical temperature.
(b) at the critical temperature.
(c) at temperature much higher than the critical temperature.
(d) at all the temperatures.
4. In Joule’s expansion of an ideal gas, the temperature of gas
(a) increases
(b) decreases
(c) remains unchanged
(d) first decreases and then increases
5. van der Waals forces are a result of interaction of
(a) only dipole-dipole
(b) only dispersion
(c) only dipole-induced dipole
(d) all of these
Real Gases, Liquefaction of Gases and Production and... 3.49

6. The critical temperature of a gas is


8a
(a) 3b (b) _____
27Rb
8a a
(c) ____ (d) ____2
27b 27b
7. The highest temperature at which a gas can be liquefied is called the
(a) Critical temperature (b) Boyle temperature
(c) Melting point (d) Boiling point
8. The critical temperature of water vapour is
(a) 0°C (b) below 0°C
(c) 100°C (d) above 100°C
9. Hydrogen gas cannot be liquefied at the room temperature because
(a) its density is very low
(b) it is a diatomic gas
(c) its thermal conductivity is quite high
(d) its critical temperature is below the room temperature
10. In Joule’s expansion of a van der Waal’s gas, the temperature of gas
(a) sometimes decreases and sometimes increases
(b) increases
(c) remains unchanged
(d) decreases
11. In Joule-Thomson expansion, the temperature of inversion of a gas for a
given pressure difference is 260°C. It can be cooled by keeping its initial
temperature at
(a) only at 260°C (b) 15 K
(c) 0°C (d) 3 K
12. The Boyle temperature TB and temperature of inversion Ti are related as
(a) Ti = TB (b) Ti = 0.5TB
(c) Ti = 2TB (d) Ti = 6.75TB
13. At Boyle’s temperature, a van der Waal gas
(a) behaves like an ideal gas
(b) cannot be liquefied
(c) causes heating effect on Joule-expansion
(d) has maximum deviation from the behaviour of an ideal gas
14. Liquefaction of helium is possible
(a) at ordinary temperature (b) below – 268°C
(c) at – 240°C (d) below – 83°C
3.50 Heat and Thermodynamics

15. Minimum temperature can be produced by


(a) adiabatic demagnetization
(b) vaporization of liquid at low pressure
(c) adiabatic expansion
(d) none of these

Answers
1. (a); 2. (d); 3. (c); 4. (b); 5. (d); 6. (b); 7. (a);
8. (d); 9. (d); 10. (d); 11. (d); 12. (c); 13. (a); 14. (b);
15. (a).
4
Chapter

The First Law of Thermodynamics

4.1 INTRODUCTION
As pointed out in the previous chapter, the first law of thermodynamics is an
extension of the principle of conservation of mechanical energy. Such an
extension became reasonable after it was shown that expenditure of work could
cause production of heat. The first quantitative experiments on inter conversion
of work and heat were carried out by Benjamin Thompson. But in 1840, the
law of conservation of energy was accepted in purely mechanical systems and
the inter-conversion of heat and work was well established (the conversion
factor was named the mechanical equivalent of heat, denoted by J). The value
of J calculated by Mayor was 3.56 J cal–1. The accepted modern value of J is
4.484 J cal–1. Mayor was able to state the principle of conservation of energy,
the first law of thermodynamics (in general terms) and to give one (rather rough)
numerical example of its application. The exact evaluation of J and the proof that
it is a constant independent of the method of measurement was accomplished
by Joule. Although Mayer was the father of the philosophy of the first law of
thermodynamics, the precise experiments carried out by Joule firmly established
the law on an experimental and inductive foundation.

4.2 WORK DEPENDS ON THE PATH INDICATOR DIAGRAM


A large mass heat source which can supply any
amount of heat to a system of finite mass without
changing its own prescribed temperature is called a
heat reservoir (Fig. 4.1).
To raise the temperature of the system by a given
amount, we can either add heat Q without performing
any work, or perform work W on the system without Fig. 4.1 A system in thermal
adding any heat such that Q = – W. (The Q is contact with a reservoir
considered positive when heat enters the system and
W is positive when work is done by the system).
Let us calculate the quantities Q and W for a simple process. Consider a gas
in a cylinder with a movable piston at one end and a heat reservoir at the other
4.2 Heat and Thermodynamics

Fig. 4.2 Energy added to the system (gas) as a result of thermal contact with the reservoir is called
heat; energy added by the displacement of piston is called work (After B.K. Agarwal “Thermal Physics”,
Lokbharti Publications, Allahabad, India (1988))

end (Fig. 4.2). The walls of the cylinder and the piston are made of insulating
(adiabatic) material. Let gas be the system. The piston and the heat reservoir form
the environment with which the system can interact.
The temperature of the reservoir can be changed in steps as desired, or we can
think of a series of reservoirs with temperatures q, q + dq, q + 2dq… etc. The heat
Q can flow into the system or out of it due to thermal contact with the reservoir
and work W can be performed on or by the system by compressing or expanding
the gas with the piston.
When any of the properties (for example, volume, pressure temperature, etc.)
of a system change, the state of the system changes and the system is said to
undergo a process. For example, when one of the weights on the piston (Fig. 4.2)
is removed, the piston rises and a change in state occurs.
The process is said to be quasi-static (i.e., almost static) when it takes place
in such a way (usually very slowly) that at every instant the system differs only
The First Law of Thermodynamics 4.3

infinitesimally from an equilibrium state. A quasistatic process can be regarded


as a succession of equilibrium states. If the weights on the piston in Fig. 4.2 are
small and are taken off one by one, the process could be considered quasistatic.
(Many actual processes closely approach a quasistatic process).
Consider a quasistatic process in which the gas expands by moving the piston
of area A against a pressure p through an infinitesimal distance dx (Fig. 4.2). The
work performed by the system is
dW = p (Adx) = pdV (1)
where we have assumed the piston to be friction-less. By our convention, the
work done by the system is positive and is consistent with the fact that the volume
change dV = + Adx in an expansion process is positive.
If a finite change in volume is carried out in such a way that the external
pressure p is known at each successive state of expansion (or compressor), we
can plot the process on a p-v diagram. Such a plot is called an indicator diagram
(Fig. 4.3). The work done by the system in the expansion process i Æ f, is
Vf

Wi, f = Ú pdV (2)


Vi

and is equal to the area under the curve. To evaluate the integral mathematically, we
have to know how pressure p varies with V (for a perfect gas pV/T a constant).

Fig. 4.3 (a) Indicator diagram for calculating work done by a gas during expansion;
(b) The work done by a system depends upon the path
4.4 Heat and Thermodynamics

It is possible to take the system from the initial state i, to the final state f in
many different ways. For example, the pressure can be kept constant from i to a
and then the volume kept constant from a to f (Fig. 4.3). Then the work done by
the gas is given by the area under the line ai. Another possible path is ibf, in which
case the work done by the gas is different and is given by the area under the line
bf. To carry out processes iaf and ibf, the temperature of the heat reservoir must
be changed during the process. The continuous path from i to f is another possible
path in which the work done by the gas is still different from the previous cases.
We conclude, therefore, that the work done by the system depends not only on the
initial and final states but also on the intermediate states i.e., on the path taken by
the process.
Clearly, it is meaningless to speak of ‘the work in (or of) a system’ in the same
way that we can speak of the pressure or temperature of a system. Work is not a
property of the system.
At this stage, we should consider point and path functions or to use another term,
exact (or complete) and inexact (or incomplete) differentials. Thermodynamic
properties like p, V or q (empirical temperature) are point functions. This means
that for a given point on the diagram (e.g., Fig. 4.3), the state is fixed, and thus
there is a definite or exact value of each property corresponding to this point. The
differentials dV of a point function V are exact differentials and can be integrated
along any path
2

Ú dV = V2 – V1
1

We can speak of volume V2 in state 2 and the volume V1 in state 1. The change
in volume V2 – V1 is independent of the path and depends only on the initial and
final states.
Work on the other hand, is a path function. The differentials dW of path
functions are inexact differentials. We generally use the symbol d to denote
inexact differentials instead of d (used for exact differentials). The integration of
dW depends on the path and so for work we write
2

Ú dW = W1, 2
1

as we cannot speak of work W2 in state 2 and W1 in state 1 and so cannot write


W2 – W1.

4.3 HEAT DEPENDS ON THE PATH


A result similar to that for work is obtained if we compute the flow of heat Q into
the system during the quasistatic process when the temperature of the reservoir
The First Law of Thermodynamics 4.5

(Fig. 4.2) is slowly changed in infinitesimal steps. The initial state i is characterized
by a temperature qi and the final state by qf .
We can heat the system (gas) at a constant pressure pi, for example, until we
reach the temperature qf , and then change the pressure to the final value pf keeping
the temperature constant. Or we can first decrease the pressure to pf and then heat
it to the final temperature qf keeping the pressure constant. Or we can follow
many other different possible paths. Each path involves a different quantity of
heat flowing into the system. This is an experimental fact. We can, therefore, say
that the heat lost or gained by a system depends not only on the initial and final
states but also on the intermediate states, that is, on the path taken by the process.
Thus, heat, like work, is a path function or an inexact differential dQ. Heat is not
a property of the system. On integrating we write
2

Ú dQ = Q1, 2
1

In words, Q1, 2 is the heat transferred during the given process between state 1
and state 2.

4.4 FIRST LAW OF THERMODYNAMICS AND


INTERNAL ENERGY
Thermodynamics can be defined as the science that deals with heat and work and
those properties of matter that bear a relation to heat and work. Like all other
sciences, the basis of thermodynamics is experimental observation.
When a system in a given state goes through a series of different processes
and finally returns to its initial state, the system is said to have undergone a cycle.
At the conclusion of a cycle all properties have the same value that had at the
beginning.
Consider a system (gas) that undergoes a cycle, changing from state i to state f
along path iaf (called process a), and returning from state f to state i along path fbi
(called process b). Such a cycle is shown, for example, in Fig. 4.3b. An alternative
possible return path is fci (called process c).
The first law of thermodynamics states that for a closed system undergoing a
cycle, the cyclic integral of the work is equal to the cyclic integral of the heat,

dW = J dQ (1)

The symbol designates the cyclic integral. If both W and Q are expressed in
the same unit, the first law is

dW = J dQ (first law for a cycle) (2)


4.6 Heat and Thermodynamics

Every experiment that has been conducted thus far, for a wide variety of systems
and for various amounts of work and heat, has verified the first law. The basis of
this law is, therefore, experimental evidence. It has never been disproved.
Often we are interested in a process rather than in a cycle. We can formulate
the first law for a closed system that undergoes a change of state. This can be done
by introducing a new property, called the internal energy U.
From the first law of thermodynamics, Eqn. (2), we have for two separate
processes a and b, Fig. 4.3(b),
(a)f (b)i (a)f (b)i
Ú dW + Ú dW = Ú dQ + Ú dQ
(a)i (b)f (a)i (b)f

If the system returns to the state i by process c, we can write


(a)f (c)i (a)f (c)i
Ú dW + Ú dW = Ú dQ + Ú dQ
(a)i (c)f (a)i (c)f

Subtracting the second of these equations from the first, we have


(c)i (b)i (c)i
(b)i
Ú (b)f dW – Ú dW = Ú dQ – Ú dQ
(c)f (b)f (c)f

or, by rearranging,
(b)i (c)i
Ú (dQ – dW) = Ú (dQ – dW)
(b)f (c)f

Since b and c represent arbitrary processes between states i and f, we conclude


that the quantity (dQ – dW) is the same for all processes between state i and state
f. Therefore, (dQ – dW) depends only on the initial and final states and not on the
path followed between the two states. It is a point function. Physicists are always
happy to discover such quantities and like to give them a name. Since both Q and
W are different forms of energy, we conclude that there exists an energy, called
the internal energy U, such that the first law reads
dU = dQ – dW (first law for a process) (3)
Since U is a point function, a property, its derivative is written dU. When
Eqn. (3) is integrated from an initial state i to the final state f, we have
Uf – Ui = Qi, f – Wi, f (4)
where Qi, f is the heat transferred to the system during the process i Æ f, Wi, f is the
work done by the system during the process i Æ f, and Ui and Uf are the initial and
final values of the property U of the system. U is an extensive property.
The first law Eqn. (4) is consistent with the fact that if we want we can add any
amount of energy to a system in the form of work (turning of paddle in Joule’s
arrangement) and take away any amount in the form of heat (by connecting the
The First Law of Thermodynamics 4.7

water to a cooler system). Therefore, neither heat Q nor work W can be conserved
alone. However, the difference Q – W = DU is conserved. The essential content of
the first law is that there exists a useful quantity called internal energy U. Internal
energy U is the energy possessed by the body which can be drawn off either as
heat or as work, depending upon the circumstances.
The first law as stated in Eqn. (4) implies the familiar principle of energy
conservation. The change in the energy of a system during a thermodynamic
process is exactly balanced by the energy exchanged with the surroundings in the
form of heat transferred to the system and in the form of work done by the system.
This justifies the three sign conventions regarding heat Q, work W and internal
energy U adopted in thermodynamics, shown in Fig. 4.4.

Fig. 4.4 Sign conventions with Q, W and U regarded as positive quantities or positive numbers

Note that in thermodynamics, the sign convention for work is opposite to that
in mechanics. The convention here is designed to fit the behaviour of heat engines
where there is an inflow of heat and an output of work. If a system receives both
heat energy and work, Eqn. (3) would read

dU = dQ – (– dW) (7)
The law of energy conservation will include the changes in the kinetic energy
(system may gain velocity) and the changes in the potential energy (system may
be elevated). Then total energy
E = internal energy + kinetic energy + potential energy
= U + KE + PE (6)
dE = dU + d(KE) + d(PE) (7)
Then, for a closed system, we can write
dE = dQ – dW (8)
4.8 Heat and Thermodynamics

4.5 WORK DONE BY A THERMODYNAMIC SYSTEM IN


EXPANSION AGAINST EXTERNAL PRESSURE
Consider a conducting cylinder fitted with a movable piston with a gas enclosed
in it. Since material of cylinder is conducting, there can be an exchange of heat
from its surroundings. Some weights can be placed on the piston. If weights
are withdrawn, gas expands and the gas does some work in expansion against
external pressure. On the other hands, if weights are increased, gas is compressed
and work is done on the gas in compression by external pressure.
Force F = Pressure × Area
=P×A (1)

B B

DX

A A
Gas pressure P
Cylinder

Fig. 4.5 Expansion of gas

If weights are withdrawn, gas expands. Let the piston moves from A to B, and
work done by the gas in expansion
DW = force × displacement
= P × A × Dx
= PDV (2)
where DV = ADx (increase in volume of the gas)
If the volume increases from V1 to V2, at a pressure P, total work done by the
gas.
V2

W = Ú PdV (3)
V1

(where dV is a small increase in volume)


If the volume decreases by DV, at a pressure P, work is done on the gas:

DW = – PDV (4)
If the volume decreases from V1 to V2, at a pressure P, total work done by the
gas
The First Law of Thermodynamics 4.9

1(i)
P

2(f)

Vi Vf
V
Fig. 4.6 PV diagram

V2

W = – Ú PdV (5)
V1

(where dV is a small increase in volume)


Equation (5) is analogous to the general result for the work done on a system
by a variable force. If we plot F against x, the work done by F is simply area under
the curve between xi and xf. Similarly, a graph can be obtained by plotting P on
the vertical axis and V on the horizontal axis. This is called a PV diagram.
The magnitude of work done on the gas is equal to the area under the curve
representing the process on a PV diagram. The sign of W is determined according
to whether Vf > Vi (in which case W is negative) or Vf < Vi (in which case W is
positive), if W represents work done on the gas.

4.5.1 The Pressure-Force is Non-Conservative


This has been demonstrated in Fig. 4.7.
We can find W1, the work done on the gas along path 1 by considering the work
done along the two segments AB and BD.

W1 = WAB + WBD

Because the volume is constant along AB, it follows that WAB = 0. Along BD,
pressure is constant at pf, it comes out of the integral and the result is
4.10 Heat and Thermodynamics

Pf B 1 D

P 1 2

Pi C
A 2

E F
Vi Vf
V
Fig. 4.7 A gas is taken from the pressure and volume at point A to the pressure and volume at point
D along two different paths ABD and ACD. Along path 1 (ABD), the work is equal to the area of rectangle
BDFE, whereas along path 2 (ACD) the work is equal to the area of rectangle ACFE

W1 = WAB + WBD

= 0 – Ú pdV
Vf

= – pf Ú dV
Vi

= – pf (Vf – Vi) (6)


For path 2 (ACD)
W2 = WAC + WCD

= – Ú pdv + 0
Vf

= – pi Ú dV
Vi

= – pi (Vf – Vi) (7)


Clearly, W1 π W2 and the work depends on the path. So, pressure force is non-
conservative.

4.5.2 Work Done at Constant Volume


The work is zero for any process in which the volume remains constant (as in
segments AB and CD in Fig. 4.7.
W = 0 (constant V)
The First Law of Thermodynamics 4.11

We deduce directly from Eqn. (5) that W = 0 if dV = 0 (or V = constant).


Note that it is not sufficient that the process start and end with the same volume;
the volume must be constant throughout the process for the work to vanish. For
example, consider process ACDB in Fig. 4.7. The volume starts and ends at Vi, but
the work is certainly not zero. The work is zero only for vertical paths such as AB,
representing a process at constant volume.

4.5.3 Work Done at Constant Pressure


Here, we can easily apply Eqn. (5), because the constant P comes out of the
integral,
W = – p Ú dv

= – P(Vf – Vi) (8)


Examples are the segments AC and BD in Fig. 4.7. Note that the work done on
the gas is negative for both of these segments, because the volume increases in
both processes.

4.5.4 Work Done at Constant Temperature


If the gas expands or contracts at constant temperature, the relationship between
P and V, given by ideal gas law (PV = nRT)is
PV = constant
On a PV diagram, the plot of the equation PV = constant is exactly like a plot
of the equation xy = constant: it is a hyperbola, as shown in Fig. 4.6.
A process done at constant temperature is called an isothermal process and the
corresponding hyperbolic curve on the PV diagram is called an isotherm.
To find the work done on a gas during an isothermal process, we use Eqn. (5),
but we must find a way of carrying out the integral when p varies. To do this, we
write p = nRT/V, then
Vf

W = – Ú pdv
Vi

Vf
nRT
= – Ú ____ dV
Vi V

Vf
dV
= – nRT Ú ___ (9)
Vi V

(because T is constant)
4.12 Heat and Thermodynamics

Carrying out the integral, we find,


Vf
W = – nRT ln __
Vi () (T = constant) (10)

Note that this is negative whenever Vf > Vi (ln x is positive for x > 1) and
positive whenever Vf < Vi.

4.5.5 Work Done in Thermal Isolation


Suppose, a gas cylinder which was initially in contact with a thermal reservoir
(below it) is now placed on a slab of insulating material (after removing the
thermal reservoir). The gas will then be in complete thermal isolation from its
surroundings, if we do work on it, its temperature will change, in contrast to its
behavior when it was in contact with the thermal reservoir. A process carried out
in thermal isolation is called an adiabatic process.
If we allow the gas to change its volume with no other constraints, then, the
path it will follow is represented on a pV diagram by the parabola-like curve

pV g = constant
as shown in Fig. 4.8.
where the dimensionless parameter g is the ratio of the specific heats Cp and Cv
(determined by experiments for any particular gas). Its values are typically in
the range 1.1 to 1.8. Because g is greater than 1, the curve pV g = constant at any

Fig. 4.8 An adiabatic process is represented on a pV diagram by the hyperbola-like curve


pV g = constant. The work done in changing the volume is equal to the area under the curve
between Vi and Vf.
The First Law of Thermodynamics 4.13

point at which they intersect (see example 1.4). As Fig. 4.8 shows, this means
that the work done by the gas in expanding adiabatically from Vi and Vf will be
somewhat smaller in magnitude than the work done in expanding isothermally
between these same two volumes.
We can find the “constant” in Eqn. (10), if we know g and also the pressure and
volume at any particular point on the curve. If we choose the initial point pi, Vi in
Fig. 4.8, the “constant” has the value piVig and we can write Eqn. (10) as
pV g = piVig (11)
g
piVi
or p = ____ (12)
Vg
We can now find the adiabatic work
Vf

W = – Ú pdV
Vi
Vf Vf
piVig dV
= – Ú ____
g
dV = – piVig Ú ___g
Vi V Vi V

piVig
= – ____ V1i – g – V1f – g
( )
g–1
By bringing a factor of V gi – 1 inside the bracket, we can write

[( ) ]
piVi Vi g–1
W = ____ __ –1 (13)
g–1 Vf
If the gas expands then Vi /Vf < 1 and since a number less than one raised to any
positive power remains less than one, the work W is negative. By further using
piVig = pfVfg (14)
we can write the adiabatic work in the equivalent form as
1
W = ____ (pfVf – piVi)(adiabatic) (15)
g–1

4.6 PROOF FOR THE FIRST LAW OF THERMODYNAMICS


The first law of thermodynamics is essentially Q W
a balance sheet of energy transactions and an
extension of the principle of conservation of DE
energy. To illustrate this point, let us consider a (increase)
closed system whose energy we want to alter.
This can be achieved through heat operation or Fig. 4.9 Operations as being done on
a closed system to alter its energy
4.14 Heat and Thermodynamics

work operation or both. If heat Q is supplied to the system and a work W is work
done on the system the consequence will be an increase in energy E of the system.
The situation is as depicted pictorially in Fig. 4.9.
The energy-balance can now be written as
Q + W = DE (increase)
= Efinal – Einitial (1)
The positive signs with Q, W and DE indicate a gain be the system. If heat is
supplied to the system and a work is done by the system, as a consequence,
Q – W = DE (2)
In the differential form, Eqn. (1) can be written as
dQ + dW = dE (3)
If internal energy U is the only energy the system has Eqn. (3) gives
dQ + dW = dU (4)
Equations (1) to (4) are quantitative statements of the first law of thermodynamics,
and may be states as follows
“The change DU in internal energy must be equal to the energy absorbed in
the process from the surroundings in the form of heat minus the energy lost to the
surroundings in the form of external work done by the system” i.e.,
DU = DQ – DW

4.7 THE THERMODYNAMIC PROPERTY ‘ENTHALPY’


In analyzing certain processes, we come
across, certain combinations of thermodynamic 2
properties which are as a result, also properties
of the substance undergoing the change of state.
For example, let us consider a control-mass*
undergoing a quasi-equilibrium constant-pressure 1
process as shown in Fig. 4.10.
Assume that there are no changes in kinetic Q Gas

or potential energy and that the only work


done during the process is that associated with
Fig. 4.10 The constant-pressure
quasi-equilibrium process

*
In case of a system with a control surface that is closed to mass flow, so that no mass can escape
or enter the control volume, it is called a control mass (containing the same amount of matter at all
times).
The First Law of Thermodynamics 4.15

the boundary movement. Taking the gas as our control mass and applying the first
law, we have (in terms of Q)

1Q2 = U2 – U1 + 1W2
2
The work done can be calculated as 1W2 = Ú PdV since the pressure is constant
1
2
\ 1W2 = P Ú dV = P(V2 – V1)
1

Therefore,
1Q2 = U2 – U1 + P2V2 – P1V1 (in general)

= (U2 + P2V2) – (U1 + P1V1)

Thus, we find that in this very restricted case, the heat transfer during the
process is given by change in the quantity U + PV between the initial and final
states. Because all these quantities are thermodynamic properties i.e., functions
only of the state of the system, their combination must also have these same
characteristics. Therefore, we find it convenient to define a new extensive property,
the enthalpy:
H ∫ U + PV
or per unit mass
h = u + Pv
As for internal energy, we could speak of specific enthalpy h, and total enthalpy
H. However, we refer to both as enthalpy, and the context will make it clear which
is being used.

4.8 HEAT CAPACITY


We define heat capacity as the amount of heat required to raise the temperature
of 1 mole of a system by 1°C. It is denoted by C. In order to define heat capacity,
two notations are used as follows:
1. Specific heat capacity: Heat absorbed by 1 g of material that undergoes a
rise in temperature of 1°C.
2. Molar heat capacity: Heat absorbed by 1 g mol of material that undergoes
a rise in temperature of 1°C.
Thus, molar heat capacity is the specific heat capacity multiplied by the gram
molecular weight.
4.16 Heat and Thermodynamics

The heat effect (Q) in a process for a change in temperature from T1 to T2 can
be calculated from the relation
Q = CDT
= C(T2 – T1) (1)
assuming that heat capacity C (which generally varies with temperature), remains
essentially constant between T1 and T2. Then heat absorbed
T2
Q = Ú (C as a function of T) dT (2)
T1

To obtain a quantity which is characteristic of a substance and which does


not depend on how much of the substance one considers, we define the specific
heat capacity c of a substance as the ratio of the true heat capacity C of a system
composed of that substance to the mass m of the system i.e.,
C
c = __
m (3i)

then, the true molar specific heat capacity would be defined as


C
c = __
n (3ii)

where n is the number of moles.

4.9 SPECIFIC HEAT CAPACITY


In the preceding section, we have talked about C as though it were a well defined
quantity but, of course, the heat required to pass from one specific state of a
system to another will vary with the path taken. Then, heat capacity will be well
defined for a specific path. In fact, we find useful two species of heat capacities,
corresponding to two simply specified paths, namely that at constant volume and
that at constant pressure.
(i) At constant volume, over a temperature range in which the heat capacity is
constant, Eqn. (1) becomes
Qv = Cv DT (4)
but Qv = DU definition, hence
DU = Cv DT (5)
If over the temperature range concerned, Cv, is not constant, then from
Eqn. (2)
T2
DU = Qv = Ú (Cv as a function of T) dT (6)
T1
The First Law of Thermodynamics 4.17

For n mol,
T2
DU = n Ú Cv dT
T1

= nCv (T2 – T1) = nCv DT (7)


if Cv is constant.
(ii) At constant pressure, we have similar to Eqn. (4)
Qp = Cp DT (8)

but Qp is by definition, the heat content or enthalpy, then


DH = Cp DT (9)
If Cp varies over the temperature range concerned, then, we have
T2

DH = Qp = Ú (Cp as a function of T) dT (10i)


T1

For n mole
T2

DH = n Ú Cp dT
T1

= nCp (T2 – T1) = nCp DT (10ii)


if Cp is constant

4.10 RELATIONSHIP BETWEEN Cp AND Cv


When an ideal gas is heated at constant volume, no work is done, and all the
heat goes into increasing the temperature of the gas. If an ideal gas is heated
at constant pressure, although the same amount of heat is still required to raise
the temperature of the gas, an additional investment of heat is required for the
PDV work done by the heated gas, as at constant pressure, it expands against the
atmosphere. Then, if Cv is the heat required to increase the temperature of 1 mol
of an ideal gas by 1°C at constant volume and Cp is the corresponding quantity at
constant pressure, we have
Cp = Cv + PDV (11)
However, from the ideal gas law, we have PV = nRT. Then for 1 mol of gas,
i.e., for n = 1, PV = RT. This can also be written (for small changes in volume and
in temperature) as PDV = RDT. By definition of “heat capacity”, the quantity of
material is 1 mol and the temperature increase is 1°C. Hence,
4.18 Heat and Thermodynamics

PDV = RDT = R(1) = R (12)


Combining Eqns. (11) and (12), we get
Cp = Cv + R (13)
This equation has been abundantly confirmed for real gases at reasonably low
pressures. The ratio of the heat capacities is given by
Cp Cv + R
g = ___ = ______ (14)
Cv Cv

4.10.1 Heat Capacities of Gases


We have seen in Chapter 2 (article 2.8) that on the basis of the Boltzmann statistics,
if the energy associated with any degree of freedom is a quadratic function of
the variable specifying the degree of freedom, the mean value of corresponding
energy-equals kT/2. For example, the kinetic energy associated with the velocity
component vx is 1/2 mvx2, and so, its mean value is kT/2. Similarly for rotation,
(kinetic energy = 1/2 Iw2), the mean rotational kinetic energy is kT/2, and for a
harmonic oscillator, where the potential energy is Kx2/2 (K = force constant),
the mean potential energy is kT/2. In general, from the principle of equipartition
of energy, the mean total energy of a molecule with f degrees of freedom, is
therefore,
_ __f
e = kT
2
and the total energy of N molecules is
_ f f
Ne = __ NkT = __ nRT
2 2
where n is the number of moles and R, the universal gas constant.

Classical Theory
In thermodynamics, the change in internal energy U of a system, between two
equilibrium states, is defined by the equation
Ua – Ub = Wad
where Wad is the work in any adiabatic process between the states (only changes
in internal energy are defined).
Starting with a molecular model of a system (e.g., monatomic, diatomic etc.),
we can identify the internal energy with the sum of the energies of the individual
molecules. Just above, we have desired a theoretical expression for the total
energy associated with the f degrees of freedom of each of the N molecules of a
gas. We thus set this equal to the internal energy U:
f f
U = __ NkT = __ nRT
2 2
The First Law of Thermodynamics 4.19

The specific internal energy per mole is


U __f
u = __
n = 2 RT
How can we test the validity of the atmosphere made in the foregoing derivation?
The most direct way is from measurements of specific heat capacities. The molal
specific heat capacity at constant volume is
u
( )
Cv = ___
∂T v
Hence, if the hypothesis above is correct, we should have

d f
( ) f
Cv = __ __ RT = __ R
dt 2 2
But for an ideal gas (from thermodynamic reasoning)
Cp = Cv + R (from Eqn. 13, Section 4.10)
f f+2
\ Cp = __ R + R = ____ R
2 2
f____
+2
Cp 2 f+2
and g = ___ = _____ = ____
Cv f__ f
2
Thus, while the principles of thermodynamics give us only an expression
for the difference between the specific heat capacities Cp and Cv, the molecular
theory together with the equipartition principle, predicts the actual magnitudes of
the specific heat capacities and their ratio g, in terms of the number of degrees of
freedom f and the universal gas constant R. Note that according to the theory, Cp,
Cv and g are all constants independent of the temperature.
Consider first a gas whose molecules are monatomic and the energy is wholly
kinetic or translational. Since, here, there are three translational degrees of
freedom, f = 3, and we expect
f 3
Cv = __ R = __ R = 1.5 R
2 2
f+2 5
Cp = ____ R = __ R = 2.5 R
2 2
Cp 5
and g = ___ = __ = 1.67
Cv 3
This is in good agreement with the values of cv and g for the monatomic gases
listed in Table 4.1.
4.20 Heat and Thermodynamics

Table 4.1 Molal specific heat capacities of a number of gases, at temperatures near room
temperature. The quantities measured experimentally are Cp and g. The former is determined
by use of a continuous flow calorimeter and the latter is obtained from measurements of the
velocity of sound in the gas.
Cp – Cv
Gas g Cp /R Cv /R _____
R

He 1.66 2.50 1.506 .991


Ne 1.64 2.50 1.52 .975
Ar 1.67 2.51 1.507 1.005
Kr 1.69 2.49 1.48 1.01
Xe 1.67 2.50 1.50 1.00
H2 1.40 3.47 2.47 1.00
O2 1.40 3.53 2.52 1.01
N2 1.40 3.50 2.51 1.00
CO 1.42 3.50 2.50 1.00
NO 1.43 3.59 2.52 1.07
Cl2 1.36 4.07 3.00 1.07
CO2 1.29 4.47 3.47 1.00
NH3 1.33 4.41 3.32 1.10
CH4 1.30 4.30 3.30 1.00
Air 1.40 3.50 2.50 1.00

Furthermore, the specific heat capacities y


of these gases are found to be practically
independent of temperature, in agreement with
the theory.
Consider next a diatomic molecule having
the dumbbell structure shown in Fig. 4.11. Its
moment of inertia about the x and y axes is very
much greater than that about the z-axis, and if
the latter can be neglected, the molecule has x
two rotational degrees of freedom. Also since
the atomic bond is not perfectly rigid, the atoms z
can vibrate along the line joining them. This Fig. 4.11 A diatomic molecule consist-
introduces two vibrational degrees of freedom ing of two atoms (considered to be point
(since the vibrational energy is partly kinetic particles) with its axis along the z-axis of
the coordinate system
and partly potential and is specified by the
separation of the atoms:
The First Law of Thermodynamics 4.21

1
2( 1
2
1
U = __ mvx2 + __ mvy2 + __ mvz2
2 )
1
( 1
2 ) (1
2
1
+ __ Ixwx2 + __ Iywy2 + __ kz2 + __ kz2
2 2 )
We might, therefore, expect seven degrees of freedom for a diatomic molecule
(3 for translation, 2 for rotation and 2 for vibration). For f = 7, the theory
predicts
7
Cv = __ R = 3.5 R,
2
9
g = __ = 1.29
7
These values are not in good agreement with those observed for the diatomic
gases listed in table. However, letting f = 5, we find
5
Cv = __ R = 2.5 R,
2
7
g = __ = 1.40
5
These are almost exactly equal to the average values of Cv and g for the diatomic
molecules in the second part of the table (except Cl2). Thus near room temperature,
these molecules behave as if either the rotational or vibrational degrees of freedom
(but not both), are shared equally with the translational degrees of freedom in the
total molecular energy.
As the number of atoms in a molecule increases, the number of degrees of
freedom can be expected to increase; and the theory predicts a decreasing ratio of
specific heat capacities, in general agreement with experiment:
(Theoretical Values):
3
Cv = __ R = 12.5 J/mol K (monatomic gas)
2
5
Cv = __ R = 20.8 J/mol K (diatomic gas)
2
Cv = 3 R = 24.9 J/mol K (polyatomic gas)

Cp = Cv + R
5
or Cp = __ R = 20.8 J/mol K (monatomic gas)
2
7
Cp = __ R = 29.1 J/mol K (diatomic gas)
2
6+2
Cp = _____ R = 4 R = 33.3 J/mol K (polyatomic gas)
2
4.22 Heat and Thermodynamics

5
g = __ = 1.67 (monatomic gas)
3
7
g = __ = 1.40 (diatomic gas)
5
4
g = __ = 1.33 (polyatomic gas)
3

4.10.2 Variation of Specific Heat of Gases with Temperature


The experimental values of cp, cv and g for monatomic gases agree with the values
obtained from the law of equipartition of energy. These values are independent
of temperature. But the values of cp, cv and g for diatomic and polyatomic gases
are found to be different at different temperatures, whereas theoretically these
values do not depend on the temperature. The variation of molar specific heat at
constant volume with temperature for hydrogen gas is shown in Fig. 4.12. It is
clear from the curve that for Hydrogen (diatomic gas), the value of cv is equal to
its theoretical value in the temperature range 250 K to 750 K. Below 250 K, the
experimental value of cv falls upto 3/2 R, while above 750 K, it rises upto 7/2 R.
This variation of specific heat with temperature could be explained by quantum
theory from which following conclusions are drawn:

3.5R

3R Vibrational

2.5R
2R Rotational
Cv 1.5R
R
Translational

0 200 K 750 K
T
Fig. 4.12 Variation of specific heat of hydrogen gas with temperature

(i) At low temperatures, a molecule of diatomic gas can have only the
translational motion (i.e., it behaves like a molecule of monatomic gas).
Therefore, each molecule has only 3 degrees of freedom. The mean energy
of a molecule is 3/2 kT and the mean energy per mole of gas is 3/2 RT.
Hence, the molar specific heat at constant volume is CV = 3/2 R.
The First Law of Thermodynamics 4.23

(ii) At intermediate (or moderate) temperatures, a molecule can have


translational motion with 3 degrees of freedom and rotational motion with
2 degrees of freedom. Therefore, the mean energy of molecule is 5/2 kT
and mean energy per mole of gas is 5/2 RT. Hence the molar specific heat
at constant volume is CV = 5/2 R.
(iii) At high temperatures, a molecule can have translational motion with 3
degrees of freedom, rotational motion with 2 degrees of freedom and
vibrational motion with 1 degree of freedom (for which mean energy is
kT). Therefore, the mean energy of a molecule is 7/2 kT, mean energy per
mole of gas is 7/2 RT and molar specific heat at constant volume is CV =
7/2 R.
The value of CV changes from 3/2 R to 5/2 R or from 5/2 R to 7/2 R at different
temperatures for different diatomic gases.
Example 4.1 A stationary mass of gas is compressed without friction from an
initial state of 0.3 m3 and 0.105 MPa to a final state of 0.15 m3 and 0.105 MPa,
the pressure remaining constant during the process. There is a transfer of 37.6 kJ
of heat from the gas during the process. How much does the internal energy of
the gas change?
Solution Q = DU + W
or Q1 – 2 = U2 – U1 + W1 – 2 (i)
V2
Here W1 – 2 = Ú pdV = p(V2 – V1)
V1

= 0.105 (0.15 – 0.30) MJ


= – 15.75 kJ
Q1 – 2 = – 37.6 kJ
Substituting in Eqn. (i), we get
– 37.6 kJ = U2 – U1 – 15.75 kJ
or U2 – U1 = – 21.85 kJ

i.e., internal energy of the gas decreases by 21.85 kJ in the process.

4.11 APPLICATIONS OF THE FIRST LAW OF


THERMODYNAMICS TO AN IDEAL GAS
4.11.1 Adiabatic Process
In an adiabatic process, the system is well is insulated so that no heat enters or
leaves, in which case Q = 0. The first law becomes in this case DU or
4.24 Heat and Thermodynamics

DEint = W (adiabatic process) (1)

Let us derive the relationship between p and V for an adiabatic process carried
out on an ideal gas. We assume the process to be carried out slowly so that the
pressure is always well defined. For an ideal gas, we can write

dEint = nCv dT (2)

(where n = number of moles and Cv is molar heat capacity at constant volume)


Thus, pdV = – dW = – dEint
= – nCv dT (3)
The equation of state of the gas can be written in differential form as
d(pV) = d(nRT)
pdV + Vdp = nRdT
or Vdp = nCv dT + nRdT
= nCv dT (using Cp – Cv = R) (4)
We now take the ratio between Eqns. (3) and (4)
nCpdT
Vdp ________
____ =
pdV – nCv dT
Cp
= – ___ ∫ – g (5)
Cv
where g denotes the ratio of molar heat capacities. Rewriting,
dp
___ dV
___
p =–g V (6)

Integrating between initial state i and final state f, we get


pf Vf
dp dV
Ú ___ ___
p -= – g VÚ V (7)
pi i

or
pf
ln __()
pi = – g ln
V
__f
Vi ()
which can be written as
piVig = pfVfg (8)
or pVg = constant (adiabatic process) (9)
We can rewrite this result in terms of temperature using the ideal gas equation
of state:
The First Law of Thermodynamics 4.25

(pV)V g – 1 = constant (10)


or TV g – 1 = constant (11)
or TiVig – 1 = TfVfg – 1 (12)

()
Vi g–1
or Tf = Ti __ (13)
Vf
Suppose we compress a gas in a adiabatic process. Then Vi > Vf. This requires
that Tf > Ti. The temperature of the gas rises as it is compressed. Conversely, the
temperature falls when a gas expands (which is often used as a means to achieve
low temperatures in the laboratory).

4.11.2 Isothermal Processes


In an isothermal process, the temperature remains constant. If the system is an
ideal gas, then the internal energy must therefore also remain constant. With
DEint = 0 the first law gives
Q + W = 0 (isothermal process, ideal gas) (14)
If an amount of (positive) work W is done on the gas, an equivalent amount of
heat Q = – W is released by the gas to the environment. None of the work done on
the gas remains with the gas as stored internal energy.

4.11.3 Constant-Volume Processed (Isochoric Process)


If the volume of a gas remains constant, it can do no work. Thus, W = 0 and the
first law gives
DEint = Q (constant volume process) (15)
In this case all the heat that enters the gas (Q > 0) is stored as internal energy
DEint > 0.

4.11.4 Cyclic Processes


In a cyclic process, we carry out a sequence of operations that eventually restores
the system to its initial state as for example, a three-step process illustrated in
Fig. 4.13. Because the process starts and finishes at point A, the change in internal
energy for the cycle is zero. Thus, according to first law

Q + W = 0 (cyclic process) (16)


where Q and W represent the totals for the cycle. In the figure, the total work
is positive because there is more positive area under the curve 3 than there is
negative area under the line 2.
4.26 Heat and Thermodynamics

Fig. 4.13 A gas undergoing a cyclical process starting at A and consisting of (1) a constant volume
process (no work), (2) a constant pressure process and (3) an isothermal process

In fact, for any cycle that is done in a counter clockwise direction, we must
have W > 0 (and therefore, Q < 0), whereas for cycles performed in clockwise
direction W < 0 (and therefore Q > 0).

4.12 IMPOSSIBILITY OF A PERPETUAL‐MOTION MACHINE


Let us imagine that a closed system undergoes a process by which it passes from
state A to state B as shown in Fig. 4.14.

System A The Process System B


UA +Q –W
UB

Fig. 4.14 A closed system undergoing a process to pass from state A to state B

If the only interaction with its surroundings is in the form of transfer of heat
Q to the system, or performance of work W by the system, the change in internal
energy U will be

DU = UB – UA = Q – W
or dU = dQ – dW
This first law is often stated in terms of the universal human experience i.e.,
it is impossible to construct a “perpetual-motion machine” viz., a machine that
produces useful work by a cyclic process with no change in its surroundings. To
see how this is embodied in the first law, consider a cyclic process from state A
to state B and back to A again. If perpetual motion were ever possible, it would
be sometimes possible to obtain net energy increase DU > 0 by such a cycle. That
this is impossible can be ascertained from the equation:
The First Law of Thermodynamics 4.27

DU = (UB – UA) + (UA – UB) = 0, (i.e., DU >| 0)


or in general

dU = 0

for any cyclic process.

4.13 SURFACES
The study of a surface, which can be considered to be a two-dimensional system
is an interesting branch pertaining to thermodynamics. The surfaces of liquids
and solids have properties distinctly different from the bulk properties of the
underlying material. These are some important examples of surfaces:
1. The interface of a liquid in equilibrium with a vapor.
2. A soap bubble stretched across a wire framework, consisting of two
surfaces trapping in between a small amount of liquid.
3. A thin film of oil on the surface of water.
A surface such as a soap bubble on a wire-framework, is somewhat like a
stretched membrane. The surface on one side of any imaginary line pulls this
line perpendicular to this line with a force equal and opposite to that exerted by
the surface on the other side of the line. The force acting perpendicular to such a
line of unit length is called the surface tension t (Greek letter ‘tau’). An adequate
thermodynamic description of a surface is given by specifying the following three
coordinates:
1. The surface tension t in newtons per meter (N/m)
2. The area of the film A (m2).
3. The absolute temperature T (K).
In dealing with a surface, which is a thin layer on a substrate, the underlying
substrate must always be considered as part of the system. This may be done,
however, without introducing the pressure and volume of the composite system,
because (as a rule), the pressure remains constant and volume changes are
negligible. The surface of a pure liquid in equilibrium with its vapour has a
particularly simple equation of state. Experiment shows that the surface tension
of such a surface does not depend on the area but is a function of the temperature
only. For most pure liquids in equilibrium with their vapour phase, the surface
tension has an equation of state of the form originated by van der Waals, but
further developed by Guggenheim:
n

( T
t = t0 1 – __
Tc ) (1)

where t0 is the surface tension at a standard temperature (usually 20°C), Tc is the


critical temperature (the temperature above which no amount of pressure will
4.28 Heat and Thermodynamics

condense vapour into liquid). It is clear from this equation that the surface tension
decreases as T increases, becoming zero when T = Tc.
The equation of state of a thin film of insoluble oil on water is particularly
interesting. If tw denotes the surface tension of a clear water surface and t the
surface tension of the water covered by the oil, then, within a restricted range of
values of area A
(t – tw) A = aT (2)
where a is a constant. The difference (t – tw) is sometimes called the surface
pressure or “two-dimensional” pressure. Equation (2) is the surface analog of the
ideal-gas law. Oil films can be compressed and expanded and when deposited on
glass, produce the interesting optical effect of Newton’s rings.

4.14 WORK IN CHANGING THE AREA OF A SURFACE FILM


Consider a soap film, which consists
of two surfaces enclosing water,
stretched across a wire framework.
L F
The right side of the wire frame-
work is movable (Fig. 4.15). If the
movable wire has a length L and the
surface tension in one surface is t,
dx
then the external force F exerted on
both surfaces is equal in magnitude Fig. 4.15 A soap film stretched across a rigid wire
framework which has a movable wire on the right.
to 2tL. For an infinitesimal An external force F displaces the movable wire an
displacement dx, the work is infinitesimal distance dx
d W = 2tL dx (3)
but for two films
2Ldx = dA
where A is the total syrface area. Hence
d W = t dA (4)
When surface tension t is expressed in newtons for meter and area A in square
meters, the work W is expressed in joules. Although Eqn. (4) has been derived for
a film with two surfaces, it is also valid for a single surface, such as the boundary
between the bulk of a liquid and the gaseous environment.
For a finite change from an initial area Ai to a final area Af,the expression for
work is
Af

W = Ú t dA
Ai
The First Law of Thermodynamics 4.29

A quasi-static process may be approximated by maintaining the external force


at all times only slightly different from the force exerted by the film.

4.15 PARAMAGNETIC ROD


A paramagnetic, unlike a ferromagnet such as iron, has no permanent magnetism
and no hysteresis. Upon being introduced into an external magnetic field, a
paramagnet becomes slightly magnetized in the direction of the field. Within the
paramagnetic material, microscopic currents are induced by the magnetic field. The
macroscopic effect of the induced currents is called the magnetic dipole moment
__› __›
per unit volume or the magnetization M. The magnetization M is analogous to the
_› __›
polarization P. Similarly, analogous to the electric displacement D, we define the
__›
magnetic field strength H.
_›
__›
B __›
H = ___ – M (1)
m0

where m0 is the permeability of vacuum a scalar quantity for isotropic paramagnets


_› _› __› __›
and B is the magnetic induction. Further, the vectors B, M and H have the same
direction.
We now introduce
__› the total magnetization M such that electrodynamic
magnetization M equals M/V multiplied by the systems volume (This is done to
make explicit the thermodynamic coordinate volume). Then, Eqn. (1) becomes
B M
H = ___ – ___ (2)
m0 V
Experiments on paramagnetic materials are not usually performed on samples
of arbitrary shape, rather, on samples in the form of a long cylinder, ellipsoid or
sphere. The restricted geometry is necessary in order to solve Ampere’s law in
closed form. We here limit ourselves to a long thin cylinder (i.e., a rod) inside
a current carrying solenoid. Most experiments on magnetic rode are performed
at constant atmospheric pressure and involve only negligible volume changes.
Consequently, we may ignore the pressure and the volume and describe a
paramagnetic solid with the aid of only three thermodynamic coordinates, which
are scalar quantities in simple thermodynamic systems:
1. The magnetic field H, measured in amperes/m (A/m).
2. The total magnetization M measured in A m2.
3. The temperature T (K)
The states of thermodynamic equilibrium of a paramagnetic solid can be
represented by an equation of state among these coordinates. Experiment shows
that the magnetization of many paramagnetic solids is a function of the ratio
4.30 Heat and Thermodynamics

of the magnetic intensity to the temperature. For small values of this ratio, the
function reduces to a very simple form, namely
Cc H
M = ____ (3)
T
which is known as Curie’s law. Cc is called the Curie constant. This SI unit for the
Curie constant in Eqn. (3) is expressed as
A.m2 K
Unit of Cc = _______ = m3.K
A/m
Since the Curie constant depends upon the amount of material, its unit may be
taken to be any one of the four listed in Table 4.2.

Table 4.2 Units of the Curie constant


Total Per mole Per kilogram Per cubic meter
3 3
m .K
____ m .K
____
m3.K K
mol kg

Materials exhibiting either electronic paramagnetism or nuclear paramagnetism


are particularly important in thermodynamics. They are used to obtain extremely
low temperatures.

4.16 MAGNETIC WORK


Consider a sample of paramagnetic material in the form of a ring of cross-
sectional area A and of mean circumference L. Suppose that an insulated wire is
wound around the sample, forming a toroidal winding of N closely spaced turns
(Fig. 4.16). A constant current may be maintained in the winding by a battery and
changed by means of the variable resistor.

Fig. 4.16
The First Law of Thermodynamics 4.31

The direct current in the windings sets up a magnetic induction B. If the


dimensions are as shown in Fig. 4.16, the B will be nearly uniform over the cross-
section of the toroid. Suppose that the current is changed and that in time dt, the
magnetic induction changes by an amount dB. Then by Faraday’s principle of
electromagnetic induction, there is induced in the winding a back emf e, where
dB
e = – NA ___ (1)
dt
During the time interval dt, a quantity of charge dz moves in the circuit and the
work done by the battery to maintain the current is given by
dW = – e dz (2)
dB
= NA ___ dz
dt
dz
= NA __ dB
dt
= NA I dB (3)
where the current I equals dz/dt.
The magnetic field H due to a current I inside a toroidal winding is given by
NI NAI NAI
H = ___ = ____ = ____ (4)
L AL V
where V is the volume of the paramagnetic material. Therefore
NAI = V H (5)
and dW = V H dB (6)
If M is the total magnetization of the paramagnetic (assumed to be isotropic),
then we have the relation
M
B = m0 H + m0 ___ (7)
V
Therefore,
dW = V H d(m0 H) + m0 H dM (8)
If no material were present within the solenoid winding then M would be zero,
B would equal m0 H and
dW = V H d(m0 H) (vacuum only) (9)
This is the work necessary to increase the magnetic field in a volume V of
empty space by an amount d(m0H). The second term m0 H dM, is the work done
in increasing the total magnetization of the material by an amount dM. We are
generally concerned with changes of temperature, energy etc. of the paramagnet
only, (brought about by work done on or by the material), consequently
4.32 Heat and Thermodynamics

dW = m0 H dM (10)
If m0 H is measured in newtons per ampere-meter and the total magnetization
M in ampere-m2 then the work W will be expressed in joules. If the total
magnetization is caused to change a finite amount from Mi to Mf , the work will
be
Mf
W = m0 Ú H dM (11)
Mi

In any actual case, a change of total magnetization is accomplished very nearly


quasi-statically and therefore, an equation of state may be used in the integration
of the expression denoting the work.

4.17 PARTIAL DERIVATIVES


In the equation of state there is a relationship between at least three variables (for
example, p, v, Q). We can choose any pair them to be independent variables; the
remaining one is then a dependent variable. We are often required to calculate the
rate of change of the dependent variable with respect to one of the independent
variables, keeping the other constant. This rate is called a partial derivative.
For example, in the case of the equation of state
Q = Q (p, v) (1)
we shall be interested in the partial derivatives

( )
∂Q
___
∂p v
and ( )
∂Q
___
∂v p
(2)

The variable which is kept constant in the process of differentiation is placed


as a subscript. This cumbersome notation is essential in thermodynamics, though
it is found superfluous in mathematics. The advantage is that the subscript v in
(∂Q/∂p)v denotes the variable v held constant and also indicates that Q is to be
expressed explicitly as a function of independent variables p and v, Q = Q(p, v).
Both of these information are important in thermodynamics.
Partial derivatives like (∂v/∂Q)p, (∂v/∂p)Q, and (∂Q/∂p)v can be interpreted
geometrically as the slopes of lines on a p-v-Q surface. For example, if AB is the
curve of intersection of a portion of an arbitrary p-v-Q surface (Fig. 4.17) with
the plane v = constant, the partial derivative (∂Q/∂p)v represents the gradient of
path AB. Its numerical value is that of the tangent of the angle between AB and
the p-v plane. The distance BB¢ is the difference in altitude which is covered by
moving along AB by dp. The principal value of this difference in altitude is given
by the product

( )
∂Q
___
∂p v
dp (3)
The First Law of Thermodynamics 4.33

Fig. 4.17 Geometrical meaning of (∂Q∂/∂p)v

The principal value of the total difference dQ in altitude, which is covered


when changes in dp and dv are considered simultaneously, or in succession in any
order, is given by the total differential

∂Q
( ) ( )
∂p v
∂Q
dQ = ___ dp + ___
∂v p
dv (4)

of the function Q (p, v)


By interchanging the roles played by the variables p, v, Q, we get two more
total differentials

( )
∂p
dp = ___
∂v Q
( )
∂p
dv + ___ dQ
∂Q v
(5)

( )
∂v
dv = ___
∂p Q
( )
∂v
dp + ___
∂Q p
dQ (6)

In general, suppose we have any three variables x, y, z, satisfying the relation


f (x, y, z) = 0 (7)
We can solve it for x and y, to get
x = x(y, z), y = y(x, z) (8)
Differentiating by parts, we have

( ) ( )
∂x ∂x
dx = ___ dy + ___ dz
∂y z ∂z y
(9)

( ) ( )
∂y ∂y
dy = ___ dx + ___ dz
∂x z ∂z x
(10)
4.34 Heat and Thermodynamics

Substituting Eqn. (10) in (9), we get

[ ( ) ( ) ] [( ) ( ) ( ) ]
∂x
1 – ___
∂y z
∂y
___
∂x z
dx =
∂x
___
∂y z
∂y
___
∂z x
∂x
+ ___
∂z y
dz (11)

This result must always be true whichever pair of variables we choose to be


independent. In particular, we can choose x and z as independent. Then, we can
give any value to dx and any other value to dz. Suppose we put dx π 0 and dz = 0.
Then (11) gives

( )( )
∂x
1 – ___
∂y z
∂y
___
∂x z
=0 (12)

or

( )
∂x
___ 1
= _______
∂y z (∂y/∂x)z
(reciprocal theorem) (13)

This is the reciprocal theorem. It allows us to replace any partial derivative by


the reciprocal of the inverted derivative with the same variable held constant.

On the other hand, if we put dx = 0 and dz π 0, we obtain

( )( ) ( )
∂x
___
∂y z
∂y
___
∂z x
∂x
+ ___
∂z y
=0 (14)

or
(∂x/∂z)y
( )
∂y
___
∂z x
= – _______
(∂x/∂y)z
(15)

We can combine Eqns. (13) and (15) to put the latter in the form

( )( ) ( )
∂x
___
∂y z
∂y
___
∂z x
∂z
+ ___
∂x y
= – 1 (reciprocity theorem) (16)

This is the reciprocity theorem. It is used to split up a derivative into a product


of more convenient derivatives.
Point functions like p, v and Q are functions of the state of the system alone.
Such functions are called state function or state variables. State variables, like
p, v, Q have specific values determined only by the equilibrium state of the system
and not by the previous history of the system. They return to the same values
whenever the system returns to the same equilibrium state. Work w and heat q,
being path functions, cannot be expressed as functions of the state of a system. It
follows that derivatives such as (dw/∂Q)p or (dq/∂p)Q have no meaning, since w
or q cannot be expressed as functions of p and Q. However, unlike w and q, the
internal energy u is a function of the state of the system and therefore its partial
derivatives will have meaning.
The First Law of Thermodynamics 4.35

For the variables p, v, Q the chain relation (16) becomes

( )( )( )
∂p
___
∂v Q
∂v
___
∂Q p
∂Q
___
∂p v
=–1 (17)

The partial derivatives constitute new physical quantities. They are all properties
and definite functions of the independent variables. Their values are fixed when
the values of the independent variables are fixed. Some of them are important
enough to be given a name.
When a system is heated at constant pressure (dp = 0), we get from Eqn. (6)
∂v
dv = ___
∂Q ( ) p
dQ (18)

If we consider the relative change in volume, we can define the coefficient of


volume expansion as

b = __( )
∂v
1 ___
v ∂Q p
(19)

When heating is performed at constant volume (dv = 0), we can use Eqn. (5) to
define the coefficient of pressure

p = __( )
∂p
1 ___
p ∂Q v
(20)

Finally, if the pressure is increased at constant temperature (dQ = 0), we can


define from Eqn. (6) the isothermal coefficient of compressibility

k = – __
v ∂p( )
∂v
1 ___
Q
(21)

The negative sign is added because for all known substances the pressure and
volume changes are of opposite signs. When these coefficients ate substituted in
(17), we get the interesting result

b=pkp (22)

4.18 CONCEPT OF A PERFECT GAS


In 1843 Joule performed the following experiment. Two copper vessels, A and B
connected by a tube with stopcock C, were immersed in a thermally isolated bath
of water, Fig. 4.18. Initially A contained air and B was evacuated.
When thermal equilibrium was attained the valve C was opened, allowing
the pressure in A and B to equalize. No change in temperature of the bath was
observed during or after this process of expansion of gas into a vacuum or free
expansion of gas.
4.36 Heat and Thermodynamics

A B

Water

Fig. 4.18 Apparatus for Joule’s experiment

As there was no change in the temperature of the bath, Joule concluded that
there had been no heat transferred to the air (system), Q = 0. In a free expansion
no work is done by the gas on its surroundings, W = 0. It is true that during the
expansion the air remaining in A is doing work on the air already in B. But this
work is done by one part of the system (air) on another part and is not done by the
system as a whole on its surroundings. The work done in free expansion cannot
be represented by an arrow piercing the system boundary.
Since Q = 0 and W = 0, Joule concluded from the first law, DU = Q – W, that
there was no change in the internal energy of the air
DU = 0 or U = constant
As the pressure and volume changed during this process, one concludes that
the internal energy of a perfect gas is independent of p and V and is a function of
temperature Q only, U = U(Q). This is called Joule’s law. Clearly the isotherms
of a perfect gas are also curves of constant internal energy.
When very careful experiments are made with real gases a very small change
in temperature is detected. Therefore, Joule’s law like Boyle’s law, holds strictly
only for a perfect gas. In fact, the perfect gas is defined as one which obeys both
Boyle’s law and Joule’s law.
In general, U is a function of any two of the variables p, V and Q. Considering
U as a function of V and Q, U = U(V, Q),

( )
∂U
dU = ___
∂V Q
∂U
( )
dV + ___ dQ
∂Q v
(1)

or, for specific internal energy,


u = u (v, Q),

( )
∂u
du = ___
∂v Q
( )
∂u
dv + ___ dQ
∂Q v
(2)
The First Law of Thermodynamics 4.37

In a free expansion of perfect gas (du = 0), there is no temperature change


(dQ = 0). Therefore

( )
∂u
___
∂v Q
=0 (perfect gas) (3)

which means u does not depend upon v.


Considering u as a function of p and Q, u = (p, Q),

( )
∂u
du = ___
∂p Q
∂u
∂Q( )
dp + ___
p
dQ (4)

For free expansion du = 0 and dQ = 0. Therefore,

( )
∂u
___
∂p Q
=0 (perfect gas) (5)

which means that u does not depend upon p. It follows that for a perfect gas u is
a function of Q only
u = u(Q) (perfect gas) (6)
We can now define a perfect gas (or an ideal gas) in thermodynamic terms as
follows
pv = RQ (7)
(definition of perfect gas)
( ) ( )
∂u
___
∂v Q
∂u
= ___
∂p Q
=0 (8)

A real gas at a pressure below about 2 atm can be regarded as a perfect gas with
only a little error in calculations. In principle, the limit of zero pressure must be
taken.
We will come across with partial derivatives again, in the next chapter.

QUESTIONS
1. Differentiate between the path function and point function.
2. State first law of thermodynamic and explain it.
3. State the first law of thermodynamics explaining the meanings of the
symbol used, hence explain (i) isothermal expansion, (ii) adiabatic
expansion and (iii) cyclic process.
4. Explain internal energy of a system on the basis of first law of
thermodynamics.
4.38 Heat and Thermodynamics

5. What is meant by P-V diagram or indicator diagram? How is the work


done on the system or the work done by the system during a change of
state obtained from this graph? Show that the work done on the system or
the work done by the system is the path function.
6. Demonstrate that the pressure-force is non-conservative.
7. Deduce expression for the work done when a perfect gas undergoes
expansion (i) adiabatically (ii) isothermally.
8. Explain “Heat depends on the path”.
9. Define the term Enthalpy. State its units and explain its physical
significance.
10. Prove that the ratio of specific heat at constant pressure to that at constant
volume for an ideal gas with f degrees of freedom per molecule is
g = Cp /Cv = 1 + 2/f. Hence show that the value of g is 1.66 for monatomic
gas, 1.4 for diatomic gas and 1.33 for polyatomic gas. How does the
experimental values agree with the theoretical values?
11. What do you mean by perpetual motion machine? Explain its physical
meaning.
12. What are surfaces. Give some examples. Derive an expression for work
required in changing the surface area of a soap film from an initial value Ai
to a final value Af.
13. Explain the concept of magnetic work by considering a sample of
paramagnetic material in the form of a ring.

OBJECTIVE TYPE QUESTIONS


1. In the perfect gas equation, PV = RT, V is the volume of
(a) 1 g gas (b) 1 mole of gas
(c) whole amount of gas (d) 1 litre gas
2. The potential energy of molecules of a perfect gas is
(a) equal to its kinetic energy (b) zero
(c) equal to its internal energy (b) infinity
3. Internal energy of a system changes in
(a) isothermal change (b) adiabatic change
(c) free expansion (d) cyclic process
4. The first law of thermodynamics is the law of conservation of
(a) momentum (b) energy
(c) both (a) and (b) (d) temperature
The First Law of Thermodynamics 4.39

5. In a cyclic process, the change in internal energy is


(a) infinite (b) zero
(c) equal to area of cycle (d) none of these

Answers
1. (b); 2. (b); 3. (b); 4. (b); 5. (b).
5
Chapter
The Second Law and Third Law
of Thermodynamics

5.1 REVERSIBLE AND IRREVERSIBLE PROCESSES


Consider a gas enclosed in the frictionless piston-cylinder arrangement, Fig. 5.1a.
We can go from the initial equilibrium state i to the final equilibrium state f in the
following three ways.

1. Irreversible Process
The piston is suddenly drawn up. Gas would rush in to fill the space, pressure
differences would be created throughout the gas volume, shock waves or sound
waves would be set up. The state of the gas is not of equilibrium and so variables
p and Q of the gas are not well defined. In fact, many variables would be required
corresponding to various values of p and Q at different points in the gas. If we
wait long enough, the thermal equilibrium with the reservoir is re-established.
This gives the final equilibrium state f, Fig. 5.1b.

Fig. 5.1 (a) Piston-cylinder arrangement; (b) irreversible process; (c) reversible process; and
(d) quasistatic irreversible process (discrete sequence of states of equilibrium)

The system passes from equilibrium state i (pi, vi) to equilibrium state f (pf , vf)
through a series of nonequilibrium states. These intermediate states, being not in
equilibrium, cannot be plotted on a p-V diagram. This is an example of a typical
irreversible process. It is not reversible because, for example, we cannot extract
the sound waves or shock waves by moving the piston down again.
5.2 Heat and Thermodynamics

2. Reversible Process
The frictionless piston is drawn very slowly, in the limit infinite slowly
(quasistatically). Then at every stage the variables p, V, Q have well defined
values because during the entire process the system remained in equilibrium.
Such a process consists of a continuous sequence of states of equilibrium and
is said to be reversible. Each successive intermediate equilibrium state can be
represented by a point in the p-V diagram. The line joining such points is called
a reversible path, Fig. 5.1(c). The process is called reversible because it can be
made to retrace its path without producing any change in the surroundings.
Reversible processes are ideal abstractions. They are never realized in actuality.
They must be carried out (i) quasistatically, and (ii) without dissipation of energy
(for example, without friction or turbulence). These conditions are never fully
satisfied.

3. Quasistatic Irreversible Process


If the process remains quasistatic but the condition of ‘no dissipation of energy’
is relaxed (for example, by slow motion of piston with friction) we no longer
have a reversible process but a quasistatic irreversible process. It follows that all
reversible processes are quasistatic, but that the converse is not true.
In a quasistatic irreversible process each successive intermediate state is a
near equilibrium state and can be represented by a discrete sequence of points,
Fig. 5.1(d).

5.2 SIGNIFICANCE OF REVERSIBLE PROCESSES


All natural process are irreversible because the two conditions of reversible
process can never be fully satisfied. If a process is irreversible it does not mean
that the system cannot be restored to its initial state. It only means that the initial
state cannot be restored ‘without producing any change in the surroundings’.
Although all natural processes are irreversible, the reversible processes
constitute convenient idealizations as they can be studied by analytical methods.
For example, consider frictionless quasistatic (i.e., reversible) expansion of a
gas. The work done by the gas is pdV. For a frictionless reversible expansion pdV
is uniquely defined for the increment dV.
If work dW is done by the system and heat dQ is added to the system, the first
law gives

dU = dQ – dW (first law) (1)


(unique) (not unique) (not unique)
The Second Law and Third Law of Thermodynamics 5.3

dU is determined uniquely by the initial and final states while dQ and dW are path
functions. However, if we specify that the work dW is a frictionless reversible
expansion, we have

dU = dQ – pdV
(unique) (unique)

It follows that dQ must also uniquely defined for this process. Therefore,

dU = dQ – pdV (2)

where each of the quantities is now in the form of a differential. If the relevant
paths are known, they can all be integrated exactly. This shows why frictionless
reversible expansions (or compressions) are so important in thermodynamics.

Some Examples of Reversible and Irreversible Processes


1. Suppose we place a hot block of metal in contact with a cool block. Then
heat is transferred from the hotter block to the cooler one. This is an
irreversible process – we cannot reverse any step in the procedure that
would cause the heat to flow in reverse direction and restore the blocks to
their original temperatures.
2. A cup of hot coffee left on a table gradually cools down. It never gets
hotter all by itself. (irreversible-process). In fact (as explained above)
all naturally occurring processes proceed in one direction only and such
spontaneous, one way processes are irreversible.
3. Consider a piece of metal on a hot plate at a temperature T. If we increase
the temperature of the hot plate by a small step dT, a small amount of
heat dQ is transferred from the hot plate to the block. If we then decrease
the temperature of the hot plate by dT, an equal amount of heat dQ is
transferred from the block to the hot plate. The block and the hot plate are
restored to their original conditions; the heat transferred in this way is by
a reversible process.
It is to be noted here that irreversibility does not imply that it is impossible
to force a process in the reverse direction but only indicates that such a reversal
cannot be achieved simply by changing parameters by infinitesimal amounts.

5.3 NEED FOR FORMULATIONS OF SECOND LAW


OF THERMODYNAMICS
Joule’s experiment for the determination of J showed that if the mechanical work
is converted into heat, there is exact equivalence.
_
W =› Q
5.4 Heat and Thermodynamics

Such a possible process a shown in Fig. 5.2(a). If, however, we wish to convert
heat into work (Fig. 5.2(b)), this is not true
_›
QπW
although admissible by the first law of thermodynamics e.g., a piece of stone
resting on the floor has never been seen to cool itself spontaneously and jump upto
the ceiling (thereby converting heat into equivalent amount of potential energy).
In all actual heat Æ work processes, we find W < Q. The purpose of the second
law is to incorporate such experimental facts into thermodynamics.

Fig. 5.2 (a) A possible process; (b) a hypothetical process

A cyclic device which would abstract heat from a single reservoir and convert
the heat completely to work is called a perpetual motion machine of the second
kind. Such a machine would not violate the first law, since it would not create
energy but would enable us (for example) to propel ships simply by extracting
heat from the oceans.
The second law rules out the possibility of constructing a perpetual motion
machine of the second kind. Thereby, it also distinguishes between the two
directions of the energy transfer viz. W Æ Q and Q Æ W.

5.4 THE SECOND LAW OF THERMODYNAMICS


From our common experience, we know that there are processes that satisfy the
law of conservation of energy yet never occur. For instance, a piece of stone
resting on the floor is never seen to cool itself spontaneously and jump upto the
ceiling. The purpose of the second law (as explained above) is to incorporate
reasons for such facts into thermodynamics. There are two equivalent statements
of the second law of thermodynamics.
Kelvin Statement There exists no thermodynamic transformation whose sole
effect is to extract a quantity of heat from a given heat reservoir and to convert it
entirely into work.
The Second Law and Third Law of Thermodynamics 5.5

Clausius Statement There exists no thermodynamic transformation whose sole


effect is to extract a quantity of heat from a colder reservoir and deliver it to a
hotter reservoir.
The meanings of the Clausius statement (C) and the Kelvin statement (K) can
be understood by referring to the impossible processes shown in Fig. 5.3. In both
statements the key word is ‘sole’. We give an example to illustrate the point.
Suppose a perfect gas expands reversibly and isothermally. Since dU = 0 in this
process, the work done by the gas is equal to the heat absorbed by the gas during
the expansion. Hence, a certain amount of heat is converted entirely into work.
However, this is not the sole effect because at the end of the process the gas
occupies a larger volume. This process is allowed by the second law.

Fig. 5.3 Schematic diagrams of the impossible processes according to (a) the Clausius Statement,
and (b) the Kelvin Statement

The two statements, C and K, are Heat Reservoir


equivalent. To prove this we show that Q1 > Q2
if K is false, and vice versa.
Proof that K false fi C false Suppose Q=W
K is false. Then we can extract heat
W=Q (Friction)
Q from a reservoir at temperature K False
Q2 and convert it entirely into work
Q
W(= Q), with no other effect. Next we
can convert this work into heat Q by Q2
means of friction and deliver it to a
Heat Reservoir
reservoir at temperature Q1 > Q2, Fig.
5.4 with no other effect. The net result
Fig. 5.4 Equivalence of Kelvin and
of this two step process is the transfer Clausius statements
5.6 Heat and Thermodynamics

of an amount of heat Q from a colder reservoir Q2 to a hotter reservoir Q1, with


no other effect. Therefore C is false.
Proof that C false fi K false First Heat Reservoir
we define an engine to be a Q1 > Q 2
thermodynamic system that can
undergo a cyclic process in which (Friction)
Q + Q¢
the system does the following things,
and only the following things: Q
Engine W = Q¢
(i) absorbs an amount of heat C False
Q1 > 0 from a hot reservoir Q
Q1;
(ii) performs an amount of work Q2

W; Heat Reservoir
(iii) rejects an amount of heat
Q2 > 0 to a cold reservoir Q2, Fig. 5.5 Equivalence of Clausius and
Kelvin statements
with Q2 < Q1.
Suppose C is false. Then we can extract Q from cold reservoir Q2 and deliver it
to hot reservoir Q1, with Q1 > Q2. Operate an engine between Q1 and Q2 for one
cycle, and arrange the engine so that it absorbs Q + Q¢ from Q1, converts Q¢ to
work W, and rejects Q to Q2, Fig. 5.5. The net result is that an amount of heat Q¢
is extracted from the single reservoir Q1 and entirely converted into work, with no
other effect. Hence K is false.

5.5 HEAT ENGINE AND ITS EFFICIENCY


Heat engine is a cyclic process in which heat is continuously converted into
mechanical work. A heat engine has the following three main parts:
(i) Source of heat It is kept at a constant high temperature T1 so that heat can
be extracted from it, without the change in its temperature.
(ii) Sink of heat It is kept at a constant high temperature T2 so that heat can be
supplied to it without the change in its temperature. The atmosphere acts
as a proper sink in practice because it has an infinite capacity to absorb
heat without change in its temperature.
(iii) Working substance It absorbs heat from the source, converts a part of it into
mechanical work and rejects the remaining heat at the sink. This is called
a cycle. Since the working substance returns back to its initial state after
a complete cycle, its internal energy* remains unchanged. Thus work is
obtained continuously due to repetition of the cycle.

* Under the same conditions of pressure, volume and temperature, the internal energy remains the
same.
The Second Law and Third Law of Thermodynamics 5.7

Figure 5.6 shows the operation of heat engine Source of Heat


T1
in one cycle. Let in a cycle, the working substance
absorbs heat Q1 from the source and rejects heat Q2
at the sink, converting the remaining energy (Q1 – Q1

Q2) into the mechanical work. Then from the first


law of thermodynamics, Q1 – Q2 = W (if there is no Cycle W
= Q1 – Q 2
loss in energy due to friction etc.).
The ratio of useful work W obtained
Efficiency of engine Q2
by the engine in one cycle to the heat absorbed Q1
Sink of Heat
from the source, is called the efficiency of engine. It T2
is written by the Greek letter h. Therefore, efficiency
of engine. Fig. 5.6 Heat engine

Work output W
h = _____________ (1)
Heat input Q1
(When W and Q1 are in same units)
Q1 – Q2 Q2
or h = _______ = 1 – ___ (2)
Q1 Q1

Obviously, the efficiency of heat engine is 1 (or 100%) only if Q2 = 0 (i.e., no


heat is rejected at the sink in a cycle). In other words, if whole amount of heat
extracted from the source is converted into the mechanical work, the engine is
said to be 100% efficient. But experimentally, we find that such an engine is not
realized in practice, i.e., no engine is 100% efficient.

5.6 CARNOT’S ENGINE: CARNOT’S CYCLE


AND ITS EFFICIENCY
To study the efficiency of an engine, Sadi Carnot in 1824, considered a theoretical
ideal engine in which there is no loss of heat due to friction etc., and the working
substance is the perfect gas. Such an engine can not be realized in practice, but is
helpful for the theoretical study as the mathematical calculations become easier.
The engine is reversible and its efficiency depends only on the temperatures of
source and sink between which it works. No engine in practice can have efficiency
more than it.
Main parts — The mainparts of the Ideal Carnot engine (Fig. 5.7) are:
(i) Source of heat,
(ii) Sink,
(iii) Perfect insulating plate,
(iv) Piston-cylinder assembly.
5.8 Heat and Thermodynamics

Fig. 5.7 Main parts of Carnot’s engine

(i) Source of heatIt is a source of heat of infinite capacity kept at a constant


high temperature T1 K, such that its temperature is unaffected on extracting
any amount of heat from it.
(ii) Sink It is kept at a constant low temperature T2 K. Its capacity is also
infinite, i.e., its temperature is unaffected even if any amount of heat is
rejected to it.
It is a perfect insulating plate which does not allow
(iii) Perfect insulating plate
heat to pass through it.
(iv) Piston-cylinder assemblyThe walls of the cylinder are perfectly insulating
and its base is perfectly conducting. Piston is also of non-conducting
material and the piston can slide inside the cylinder without any friction.
Cylinder contains the working substance (i.e., perfect gas). The cylinder
can be kept in contact with either of the heat source, insulating plate and
sink. No work is needed in taking it from one to the other.
Carnot Cycle According to Carnot, to obtain
Isoth
work continuously from the Ideal engine, it A(T1) e
expa rmal
has to be operated in a definite cycle such nsion
B(T1)
com

that the working substance returns back to


Adi ressio
aba

its initial state after each cycle. A cycle is


p

Adia sion
expa
tic

completed by the four operations in order.


batic
n

Isoth
n

These operations are represented on the P-V D comp ermal


P ressio
diagram in Fig. 5.8. The cycle is called the (T2) n
Carnot cycle. Let P1, V1 and T1 be the initial C(T2)
pressure, volume and temperature of the
gas. The four operations are:
a d b c
(i) Isothermal expansion, v
(ii) Adiabatic expansion, Fig. 5.8 Carnot cycle
(iii) Isothermal compression,
(iv) Adiabatic compression
The Second Law and Third Law of Thermodynamics 5.9

(i) First operation: isothermal expansion The base of the cylinder is kept in
contact with the heat source at temperature T1 K and the piston is slowly
withdrawn out of the cylinder so that the gas expands isothermally. During
this process work is done by the gas, but its internal energy remains constant
(since temperature remains constant = T1), therefore, the gas absorbs
heat from the source. The state of the gas changes from A (P1, V1, T1)
to B (P2, V2, T1) as shown by the curve AB on the P-V diagram. In the
process, work done by the gas W1 is equal to the heat absorbed Q1 from
the source (since there is no change in internal energy of gas).
Heat absorbed from the source Q1 = work done by the gas
V2

W1 = Ú PdV
V1
V2
RT1
or Q1 = W1 = Ú ____ dV
V1 V

V2
= RT1 loge __ (since PV = RT1)
V1
= area ABbaA enclosed by the curve AB on the
P-V diagram with the V-axis (3)
(ii) Second operation: Adiabatic expansion Now the cylinder is taken away
from the source and its base is kept in contact with the insulating plate.
Piston continues to move outwards due to inertia, i.e., the gas expands
adiabatically (since the gas is enclosed from all sides inside the insulator).
In this process, work is done by the gas due to which internal energy of gas
and hence the temperature of gas decreases from T1 to T2 (= temperature
of sink). The state of the gas changes from B(P2, V2, T1) to C(P3, V3, T2) as
shown by the curve BC on the P-V diagram.
Decrease in internal energy of gas = work done by the gas
V3
W2 = Ú PdV
V2

V3
K
or W2 = Ú __g dV
V2 V

K 1
(
= ______ _____
1
– _____
(g – 1) Vg – 1 Vg – 1
2 3
) (since P2 V2g = P3 V3g = K)

1
or W2 = ______ (P2 V2 – P3 V3)
(g – 1)
5.10 Heat and Thermodynamics

R
= ______ (T1 – T2) (since P2 V2 = RT1 and P3 V3 = RT2)
(g – 1)
= area BCcbB enclosed by the curve BC on the
P-V diagram with V-axis (4)
(iii) Third operation: isothermal compression Now the cylinder is withdrawn from
the insulating plate and is kept in contact with the sink. The piston is moved
slowly inwards the cylinder so that the gas is compressed isothermally at
temperature T2. The work is done on the gas, but the internal energy of gas
remains constant (since temperature of gas does not change), therefore,
heat is rejected at the sink. The state of the gas changes from C(P3, V3, T2)
to D(P4, V4, T2) which is shown by the curve CD on the P-V diagram.
Heat rejected at the sink Q2 = work done by the gas
V4
W3 = – Ú PdV
V3
V
4
RT2
or Q2 = W3 = Ú ____ dV
V3 V

V3
= RT2 loge ___ (since PV = RT2)
V4
= area CDdcC enclosed by the curve CD on the
P-V diagram with the volume axis (5)
(iv) Fourth operation: adiabatic compression The cylinder is now removed from
the sink and is again kept in contact with the insulating plate. This piston
is moved inwards the cylinder to compress the gas till the gas come back
to its initial state (P1, V1, T1). In this operation, the compression of gas is
adiabatic. The work is done on the gas due to which the internal energy
of gas increases and hence the temperature of gas rises from T2 to T1.
The state of gas changes from D(P4, V4, T2) to A (P1, V1, T1) which is
represented by the curve DA on the P-V diagram.
Increase in internal energy of gas = work done by the gas
V1
W4 = – Ú PdV
V4

V1
K
or W4 = – Ú __g dV
V4 V

K 1
(
= ______ _____
1
– _____
(g – 1) Vg – 1 Vg – 1
1 4
) (since P1 V1g = P4 V4g = K)
The Second Law and Third Law of Thermodynamics 5.11

1
= ______ (P1 V1 – P4 V4)
(g – 1)
R
= ______ (T1 – T2) (since P1 V1 = RT1 and P4 V4 = RT2)
(g – 1)
= area DAadD enclosed by the curve DA on the
P-V diagram with V-axis (6)
The cycle is repeated to get the work continuously.
From Eqns. (4) and (6), it is clear that W2 = W4, i.e., the work done by
the gas in adiabatic expansion = the work done on the gas in adiabatic
compression.
Hence the area BCcbB and area DAadD on the P-V diagram are equal.
The net work done by the working substance in the
Efficiency of Carnot Engine
complete cycle of Carnot engine is
W = W1 + W2 – W3 – W4
= W1 – W 3 (since W2 = W4)
= Q1 – Q2 = area ABCDA on the P-V diagram.
Putting the values of W1 and W3 from Eqns. (3) and (5)
V2 V3
W = RT1 loge ___ – RT2 loge ___ (7)
V1 V4
Since the points B and C are on the adiabatic curve BC, therefore, from
TVg – 1 = constant, we get
T1 Vg2 – 1 = T2 Vg3 – 1
or (T1/T2) = (V3/V2)g – 1 (8)

Similarly, the points A and D are on an adiabatic curve AD, therefore,


T1 Vg1 – 1 = T2 Vg4 – 1
or (T1/T2) = (V4/V1)g – 1 (9)
Form Eqns. (8) and (9), we get
V3 ___
___ V4
= =r
V2 V1
where r is the adiabatic expansion ratio.
V2 ___
___ V3
or = =r
V1 V4
where r is the isothermal expansion ratio.
5.12 Heat and Thermodynamics

\ From Eqn. (7),


W = RT1 loge r – RT2 loge r
= R(T1 – T2) loge r (10)
Heat absorbed by the working substance in one cycle
V2
Q1 = RT1 loge ___
V1
= RT1 loge r [from Eqn. (3)]
Net work output
\ Efficiency of engine h = ______________________ (11)
Heat absorbed from source

W Q1 – Q2 Q2
= ___ = _______ = 1 – ___ (12)
Q1 Q1 Q1

R(T1 – T2) loge r


or h = ______________
RT1 loge r

T1 – T2 T2
= ______ = 1 – __ (13)
T1 T1
It is clear from the above expression that
(i) The efficiency of Carnot engine depends only on the temperatures of the
source and sink. The efficiency of all the engines working between the
same two temperatures is the same.
(ii) The efficiency of an engine can be 1.0 (or 100%) only if the temperature
of sink is absolute zero. But it is impossible to acquire the temperature of
sink equal to absolute zero, hence no engine can be 100% efficient.
(iii) The efficiency of engine is more if (T1 – T2) is more.
The efficiency of the Carnot engine can be increased
Ways of Increasing the Efficiency
in two ways: (a) by keeping the temperature of source T1 constant and decreasing
the temperature of sink T2, (b) by keeping the temperature of sink T2 constant and
increasing the temperature of source T1. But out of these, the first way is more
effective because T2 < T1, therefore, the decrease in T2 is more effective than the
increase in T1 by the same amount.
Note
1. The efficiency of Carnot engine can also be expressed in terms of adiabatic
expansion ratio r. From Eqns. (8) and (9),
T1
__ = rg – 1
T2
The Second Law and Third Law of Thermodynamics 5.13

\ Efficiency of engine
T2 g–1

T1
1
h = 1 – __ = 1 – __
r () (14)

2. Comparing Eqns. (12) and (13), for an ideal engine

Q2 __
___ T2
= (15)
Q1 T1

Thus, the ratio of heat rejected at the sink to the heat absorbed from the source
is equal to the ratio of temperature of sink to the temperature of source.

5.7 WORKING OF A HEAT ENGINE/REFRIGERATOR


OPERATING ON A CARNOT CYCLE
Let us assume that we have a heat engine, operating between the given high
temperature and low temperature reservoirs does so in a cycle in which every
process is reversible. (It is also assumed that the high temperatures and the low
temperature of the two reservoirs remain constant regardless of the amount of
heat transferred). If every process is reversible, the cycle is also reversible, and if
the cycle is reversed, the heat engine becomes a refrigerator. We now consider the
heat engine (refrigerator) that operates on a Carnot cycle (i.e., the most efficient
cycle we can have) as depicted in Fig. 5.9. The figure shows a power plant that is
similar in many ways to a simple steam power plant.

High-temperature reservoir

QH QH

Boiler
(condenser)

W
Pump Turbine
(turbine) (pump)
W
Condenser
(evaporator)

QL QL

Low-temperature reservoir

Fig. 5.9 A heat engine that operates on a Carnot cycle (The refrigerator is shown by
dotted lines and parentheses)
5.14 Heat and Thermodynamics

Consider the working fluid to be a pure substance, such as steam. Heat is


transferred from the high temperature reservoir to the water (steam) in the boiler.
For this process to be a reversible heat transfer, the temperature of the water
(steam) must be only infinitesimally lower than the temperature of the reservoir.
This result also implies (since the temperature of the reservoir remains constant)
that the temperature of the water must remain constant. Therefore, the first process
in the Carnot cycle is a reversible isothermal process in which heat is transferred
from the high temperature reservoir to the working fluid.
The next process occurs in the turbine without heat transfer and is therefore,
adiabatic-process during which the temperature of the working fluid decreases
from the temperature of the high temperature reservoir to the temperature of the
low temperature reservoir.
In the next process, heat is rejected from the working fluid to the low temperature
reservoir. This must be a reversible isothermal process in which the temperature
of the working fluid is infinitesimally higher than that of the low temperature
reservoir. During this isothermal process some of the steam is condensed.
The final process, which completes the cycle, is a reversible adiabatic process
in which the temperature of the working fluid increases from the low temperature
to the high temperature.

5.7.1 Efficiency of Refrigerators


In case of a refrigerator, we supply Hot
work in order to extract an amount Q2 of T1
heat from a cold body (e.g., something
to be freezed in a refrigerator) and Q1 = (W + Q2)
then throw it out into the surrounding
(atmosphere). The heat rejected is not
just Q2 but Q1 = Q2 + W (Fig. 5.10). W
In this case, one does not define
efficiency but a coefficient of
performance: Q2

Q2
K = ___ T2
W
Cold
T2
= ______ Fig. 5.10 Schematic figure of a refrigerator. Work
T1 – T2 is done to remove heat from a cold source and then
dump it into a hot reservoir
Thus, it is obvious that K can be > 1.
For an ideal refrigerator, the variation
of K as a function of the ratio (T2/T1) is shown in Fig. 5.11. For a domestic
refrigerator K ~ 9. Figure 5.11 also reveals that K Æ 0 as T2 Æ 0. In other words,
The Second Law and Third Law of Thermodynamics 5.15

Fig. 5.11 Temperature variation of K, the coefficient of performance of an ideal refrigerator, i.e., a
Carnot engine reversed. When T2 = T1, the cold and the hot reservoirs have the same temperature.
Usually, T1 is the room temperature (~ 300 K) while T2 denotes the temperature to be attained by
cooling. Down to (T2/T1) = 0.5, more heat is absorbed than the work supplied but below that the
work required to extract a given quantity of heat increases sharply (W = Q/K). After G. Venkataraman,
“A Hot Story”, Universities Press (India) 100. (1993)]

refrigerators meant to cool objects down to T ~ 0 K are very efficient. However, if


going very low in temperature is our objective, then there is no choice.

5.8 DEMONSTRATION OF THE FACT THAT THE CARNOT


CYCLE IS THE MOST EFFICIENT CYCLE OPERATING
BETWEEN TWO FIXED‐TEMPERATURE RESERVOIRS
There are two propositions regarding the efficiency of a Carnot cycle:

First Proposition
“It is impossible to construct an engine that operates between two given reservoirs
and is more efficient than a reversible engine operating between the same two
reservoirs”.
The proof of this statement is accomplished through a “thought experiment”:
We make an initial assumption, then show that this leads to impossible conclusion.
Then, the only possible conclusion is that our initial assumption was incorrect.
This is as follows.
Let us assume that there is an irreversible engine operating between two given
reservoirs that has a greater efficiency than a reversible engine operating between
the same two reservoirs.
5.16 Heat and Thermodynamics

Let the heat transfer to the irreversible engine be QH, the heat rejected be Q¢L
and the work performed be WIE (= QH – Q¢L) as shown in Fig. 5.12.

System boundary

High-temperature reservoir

QH QH

WIE = QH – Q ¢L Irreversible Reversible


Wnet = QH – Q ¢L
engine engine

WRE = QH – Q ¢L

Q ¢L QL

Low-temperature reservoir

Fig. 5.12 [After: R.E. Sonntag, C. Borgnakke and G.J.V. Boyle Fundamentals of Thermodynamics,
John Wiley (2002)]

Let the reversible engine operate as a refrigerator (this is possible since it is


reversible). Finally, let the heat transfer with the low temperature reservoir be QL,
the heat transfer with the high temperature reservoir be QH and the work required
be WRE (= QH – QL).
Since the initial assumption was that the irreversible engine is more efficient,
it follows (because QH is the same for both engines) that Q¢L < QL and WIE >
WRE. Now, the irreversible engine can drive the reversible engine and still deliver
the net work Wnet = WIE – WRE = QL – Q¢L. If we consider the two engines and
the high temperature reservoir as a system (as indicated in Fig. 5.12), we have a
system that operates in a cycle, exchanges heat with a single reservoir and does a
certain amount of work. However, this would constitute a violation of the second
law and we conclude that our initial assumption (that the irreversible engine is
more efficient than a reversible engine) is wrong. Therefore, we cannot have
an irreversible engine that is more efficient than a reversible engine operating
between the same two reservoirs.

Second Proposition
“All engines that operate on the Carnot cycle between two given constant-
temperature reservoirs have the same efficiency”.
The proof of this proposition is similar to the proof just outlined (above),
which assumes that there is one Carnot cycle that is more efficient than the
another Carnot cycle operating between the same reservoirs. Let the Carnot cycle
The Second Law and Third Law of Thermodynamics 5.17

with the higher efficiency replace the irreversible cycle of the previous argument
and let the Carnot cycle with the lower efficiency operate as a refrigerator. The
proof proceeds with the same line of reasoning as above and details are left as an
exercise for the reader.

5.9 THERMODYNAMIC TEMPERATURE SCALE


In the first chapter, we pointed out that the zeroth law of thermodynamics provides
a basis for temperature measurement, but that a temperature scale must be defined
in terms of a particular thermometer substance and device. A temperature scale
that is independent of any particular substance, which might be called an absolute
temperature scale, would be most desirable. In the previous section, we have noted
that the efficiency of a Carnot cycle is independent of the working substance and
depends only on the temperature. This fact provides the basis for such an absolute
temperature scale which we call the thermodynamic scale.
The concept of this temperature scale may be developed with the help of
Fig. 5.13, which shows three reservoirs and three engines that operate on the
Carnot cycle. T1 is the highest temperature, T3 is the lowest temperature and
T2 is an intermediate temperature, and the engines operate between the various
reservoirs as indicated. Q1 is the same for both A and C and since, we are dealing
with reversible cycles, Q3 is the same for both B and C.

T1

Q1
Q1
WA A

C WC
Q2
T2
Q2
Q3
WB B

Q3
T3

Fig. 5.13

Since the efficiency of a Carnot cycle is a function only of the temperature, we


can write
QL
hthermal = 1 – ___ = 1 – y (TL, TH) (1)
QH
5.18 Heat and Thermodynamics

where y designates a functional relation. Let us apply this functional relation to


the three Carnot cycles of Fig. 5.13.
Q1
___ = y(T1, T2) (2)
Q2
Q2
___ = y(T2, T3) (3)
Q3
Q1
___ = y(T1, T3) (4)
Q3
Since
Q1 _____
___ Q1 Q2
= ,
Q3 Q2 Q3
it follows that
y(T1, T3) = y(T1, T2) × y(T2, T3) (5)

Note that the LHS is a function of T1 and T3 (and not of T2) and therefore,
the RHS must also be a function of T1 and T3 (and not of T2). From this fact we
conclude that the form of the function y must be such that
f(T1)
y(T1, T2) = ____ (6)
f(T2)

f(T2)
y(T2, T3) = ____ (7)
f(T3)

so that f(T2) will cancel from the product of y(T1, T2) and y(T2, T3). Therefore,
we conclude that
Q1
___ f(T1)
= y(T1, T3) = ____ (8)
Q3 f(T3)
In general terms
QH _____
___ f(TH)
= (9)
QL f(TL)
Now, there are several functional relations that will satisfy this equation. For
the thermodynamic scale of temperature, which was originally proposed by Lord
Kelvin, the selected relation is
QH ____
___ (tH)
= (10)
QL (tL)
(where tH and tL are the Kelvin temperature itself corresponding to f(TH) and
f(TL) respectively).
The Second Law and Third Law of Thermodynamics 5.19

With absolute temperature t so defined, the efficiency of a Carnot cycle may


be expressed in terms of the absolute temperatures:
QL tL
hthermal = 1 – ___ = 1 – ___ (11)
QH tH
This means that if the thermal efficiency of a Carnot cycle operating between
two given constant-temperature reservoirs is known, the ratio of the two absolute
temperatures is also known.

5.9.1 Zero of the Thermodynamic Scale


QL tL
h = 1 – ___ = 1 – __ (12)
QH tH
If tL = 0 then QL = 0 and h = 1.
Thus, zero of the thermodynamic scale is the temperature of sink at which the
reversible engine is 100% efficient (i.e., the engine converts whole of the heat into
work without rejecting any heat at the sink). This temperature is called absolute
zero. It may be mentioned here that it is not possible to define absolute zero
by any empirical gas scale because each gas becomes solid before reaching this
temperature and the thermometer does not work.
Now, we can see that zero on the thermodynamic scale is the lowest possible
temperature on the absolute scale and there is no negative temperature on this
scale.
For this, consider that a reversible engine rejects heat QL at the sink temperature
– tL, then, its efficiency
QL tL tL
QH ( )
h = 1 – ___ = 1 – – ___ = 1 + ___
tH tH

i.e., h > 1 or efficiency obtained from the reversible engine is more than
100% which is impossible. Hence, negative temperature is not possible on the
thermodynamic scale.

5.9.2 Size of Degree on the Thermodynamic Scale


To fix the size of degree, like the Celsius scale, the interval between the steam
point ts and the ice point ti is divided into 100 equal parts. Each part is called a
degree. Thus, ts – ti = 100 Kelvin degrees.
Now, if a reversible engine absorbs heat Qs at the steam point ts and rejects
heat Q at a temperature t, then
Qs __
___ ts
= (13)
Q t
5.20 Heat and Thermodynamics

Similarly, if another reversible engine absorbs heat Q at a temperature t and


rejects heat Qi at the ice point ti (where ti < t), then
Q __
___ t
=
Qi ti
Qi __
___ ti
or = (14)
Q t
From the above two equations

ts – ti
Qs – Qi _____
______ = (15)
Q t

But ts – ti = 100, therefore

Qs – Qi ____
______ 100
=
Q t

or (
Q
t = 100 × ______
Qs – Qi ) (16)

Thus, an unknown temperature t can be calculated.

5.9.3 Equivalence of the Thermodynamic Scale


with the Perfect Gas Scale
Let the temperatures of source and sink on the perfect gas scale be T1 and T2
respectively (T1 > T2). A reversible engine working between these temperatures
absorbs heat Q1 from the source and rejects heat Q2 to the sink. Then, the efficiency
of Carnot engine
Q2 t2
h = 1 – ___ = 1 – __
Q1 t1

Q2 __
___ T2
or = (17)
Q1 T1
Now, if the temperatures of these source and sink on the thermodynamic (i.e.,
absolute) scale are T1 and T2 respectively, then
Q2 __
___ t2
= (18)
Q1 t 1

Thus,
T2 __
__ t2
= (i.e., the ratios are equal) (19)
T1 t 1
The Second Law and Third Law of Thermodynamics 5.21

Now, if the engine works between the steam point and the ice point, then
Ti __
__ ti
= (20)
Ts t s

Ti ti
or 1 – __ = 1 – __
Ts ts
ts – ti
Ts – Ti _____
______
or =
Ts ts

100 ____
____ 100
or =
Ts ts

i.e., Ts = ts (21)
Similarly,
Ts __
__ ts
=
Ti ti
Ts
__ ts
fi – 1 = __ – 1
Ti ti

ts – ti
Ts – Ti _____
______
or =
Ti ti

100 ____
____ 100
or =
Ti ti

or Ti = ti (22)
i.e., the steam point and the ice point are same on the two scales. Now, if a
reversible engine operates between the steam point and any other temperature (T
or t), then
T __
__ t
=
Ts t s
\ T=t (\ Ts = ts),
i.e., both the scales are perfectly equivalent.

5.10 RELATION BETWEEN THERMODYNAMIC SCALE,


CELSIUS SCALE, RANKINE SCALE AND THE
FAHRENHEIT SCALE OF TEMPERATURES
Suppose we have a heat engine operating on the Carnot cycle that receives heat
at the temperature of steam point and rejects heat at the temperature of ice point.
5.22 Heat and Thermodynamics

To gain an idea about the magnitude of a degree on the thermodynamic scale, we


cannot construct an engine on the Carnot cycle and perform the experiment for
its efficiency. However, we can follow the reasoning as the following “thought
experiment”. If the efficiency of such an engine could be measured, we would
find it to be 26.8%. then

TL
h = 1 – ___
TH

Tice
= 1 – _____ = 0.2680
Tsteam
Tice
_____
\ = 1 – 0.2680 = 0.7320 (1)
Tsteam
This is one equation concerning the two unknowns TH and TL. The second
equation comes from deciding the magnitude of the degree on the thermodynamic
(i.e., absolute) scale to be equal to the magnitude of the degree on the Celsius
scale. Then
Tsteam – Tice = 100 (2)

Solving Eqns. (1) and (2) simultaneously, we find


Tsteam = 373.15 K,

Tice = 273.15 K (3)


It follows that
T(°C) + 273.15 = T(K) (4)
The absolute scale related to the Fahrenheit scale is the Rankine scale,
designated by R. On both these scales, there are 180 degrees between the ice point
and the steam point. Therefore, for a Carnot cycle heat engine operating between
the steam point and the ice point, we would have the two relations:
Tsteam – Tice = 180 (5)

Tice
_____ = 0.7320
Tsteam
solving these two equations simultaneously, we find
Tsteam = 671.67 R,

Tice = 491.67 R (6)


The Second Law and Third Law of Thermodynamics 5.23

It follows that since 0° R = 32°F, we have the following relation between


Fahrenheit and Rankine scales

491.67 – 32 = T (R) – T (F)

or T (F) + 459.67 = T (R) (7)

As already pointed out, the measurement of efficiencies of Carnot cycles is


not a practical way to approach the problem of temperature measurements on
the thermodynamic scale. Rather, the actual approach is based on ideal gas
thermometer and an assigned value for the triple point of water (as described in
the first chapter). At the tenth conference on Weights and Measures (held in 1954),
the temperature of the triple point of water was assigned the value 273.16 K (the
triple point of water is approximately 0.01°C above the ice point). The ice point
is defined, as the temperature of a mixture of ice and water at a pressure of 1
atmosphere (101.13 kPa) of air that is saturated with water vapour.

5.11 THE CLAUSIUS THEOREM


Consider a smooth closed curve representing a reversible cycle CR (Fig. 5.14).
Let the curve be divided into a large number of strips by means of reversible
adiabatics. Each strip may be closed at the top and bottom by reversible isotherms.
The original closed cycle is thus replaced by a zig-zag closed path consisting of
alternate adiabatic and isothermal processes such that the heat transferred during
all the isothermal processes is equal to the heat transferred in the original cycle (If
the adiabatics are close to one another and the number of Carnot cycles is large,
the zig-zag closed path will coincide with the original cycle).

P
Reversible
adiabatics

dQ3
T3
e f
T1 b Reversible
a isotherms
dQ1

T2 c g Original
d T4 reversible cycle

dQ2 dQ
4

Fig. 5.14 A reversible cycle divided into a large number of Carnot cycles
5.24 Heat and Thermodynamics

For the element cycle abcd, dQ1 heat is absorbed reversibly at T1 and dQ2 heat
is rejected reversibly at T2.*
So,
dQ1
____ dQ2
= – ____ (1)
T1 T2
(taking heat supplied as positive and heat rejected as negative)
dQ1 ____
____ dQ2
or + =0 (2)
T1 T2
Similarly, for the elemental cycle efgh
dQ3 ____
____ dQ4
+ =0 (3)
T3 T4
If similar equations are written for all the elemental Carnot cycles, then for the
whole original cycle CR
dQ1 ____
____ dQ3
dQ2 ____
+ + +…=0
T1 T2 T3
dQ
___
or =0 (4)
T CR

for any closed reversible cycle CR.


This result is known as Clausius theorem.

5.12 THE CLAUSIUS INEQUALITY


Let us consider a cycle ABCD (Fig. 5.15). Let AB be a general process, either
reversible or irreversible, while the other processes in the cycle are reversible. Let
the cycle be divided from a number of elementary cycles as shown. For one of
these elementary cycles:
dQ2
h = 1 – ____ (5)
dQ
where dQ is the heat supplied at T and dQ2 is the heat rejected at T2.
Now, the efficiency of a general cycle will be equal to or less than the efficiency
of a reversible cycle, i.e.,

( dQ2
)( ) dQ2
1 – ____ £ 1 – ____
dQ dQ rev
(6)

or
dQ2
____
dQ ( )dQ2
≥ ____
dQ rev

* Note that an imperfect differential has been represented by dQ or dQ.


The Second Law and Third Law of Thermodynamics 5.25

T=C

B
A
T dQ Reversible
adiabatics

D C
T2
dQ2

Fig. 5.15 Pertaining to Clausius inequality

or
dQ
____
dQ2 ( )
dQ
£ ____
dQ2 rev
(7)

Since ( ) dQ
____
dQ2 rev
T
= __
T 2
(8)

Therefore,
dQ __
____ T
£
dQ2 T2

dQ2
dQ ____
___
or £ for any process AB (9)
T T2
(reversible or irreversible)
For a reversible process
dQrev dQ2
ds = _____ = ____ (10)
T T2
Hence, for any process AB
dQ
___ £ ds (11)
T
and for any cycle
dQ
___ £ ds (12)
T
5.26 Heat and Thermodynamics

Since ds = 0
therefore
dQ
___ £0 (13)
T
This is known as Clausius inequality. It provides the criterion of the reversibility
of a cycle.
dQ
___
If £ 0, the cycle is reversible (14)
T
dQ
___ < 0, the cycle is irreversible and possible (15)
T
dQ
___ > 0, the cycle is impossible, since it violates the second law. (16)
T

5.13 ENTROPY B

Consider an arbitrary reversible cycle


Path I
represented by the closed curve in Fig. 5.16. p
A and B areany two points on the curve. From Path II
Clausius theorem,
we have
B A
dQ(rev) dQ(rev) dQ(rev) A
______ = Ú ______ + Ú ______
T A T B T V
(path I) (path II)
Fig. 5.16 An arbitrary reversible cycle
=0 (1)
The cycle being reversible, we can traverse the path II from B to A in the
opposite direction. Then for path II
A B
dQ(rev) dQ(rev)
Ú ______ =– Ú ______ (2)
B T A T
(path II) (path II)
So, Eqn. (1) becomes
B B
dQ(rev) dQ(rev)
Ú ______ = Ú ______ (3)
A T A T
(path I) (path II)
Since paths I and II were chosen at random, and represent any two reversible
paths, we have the result that
B
dQ(rev)
Ú ______
T
= independent of path between A and B.
A
The Second Law and Third Law of Thermodynamics 5.27

In mathematical terms, it means that the integrand dQ(rev)/T can be expressed


as the perfect differential of a point function or a state variable. Denoting this
variable by S, we can write
dQ(rev)
dS = ______ (4)
T
The point function S was introduced by Clausius in 1850, and is called the
entropy.
B B
dQ(rev)
\ SB – SA = Ú dS = Ú ______ (5)
A A T
Equation (4) indicates that when the inexact differential dQ(rev) is multiplied
by 1/T, it becomes an exact differential; 1/T is the integrating factor. The integral
B B
dQ(rev)
Ú dQ(rev) is dependent on the path whereas Ú ______
T
is independent of the path.
A A
Entropy is an extensive property. Common units for entropy are Joule/K or
cal/K.
The Clausius theorem now reads
dS = 0 (6)
Since S is a state variable, this result also follows from Eqn. (5). Entropy like
other state variables, is associated with a state and not with a process. Given
two equilibrium states A and B, it is possible to determine the entropy difference
SB – SA, regardless of the actual process (reversible or irreversible), which the
system may have undergone. To calculate this difference, we can choose any
reversible path joining them without reference to the process being analyzed (In
practice, a reversible path yielding the simplest form of the integral is chosen).
To summarize, we have the following consequences of the second law of
thermodynamics:
1. For all substances, there exists a thermodynamic temperature scale T.
2. The thermodynamic temperature T constitutes the integrating denominator
for dQ.
3. There exists a useful state variable S, called entropy.

5.14 PHYSICAL SIGNIFICANCE OF ENTROPY


Entropy is a mathematical function, which is a definite single valued function of
the thermodynamic coordinates, defining the state of a system. The entropy of a
system is a real physical quantity given by
A
dQ
SA = SO + Ú ___
O T
5.28 Heat and Thermodynamics

where SO denotes the entropy of the system in state O. Here SA is a function of


A only since the state O is fixed. The entropy thus defined, requires the arbitrary
choice of the standard state O. Thus, SA can be calculated except for an additive
constant. We also know that the entropy of a system increases in an irreversible
process and remains constant in the limit of reversible process. Thus, the flow of
heat always takes place in a direction in which the entropy increases.
It is rather very difficult to visualize the concept of entropy because there is
nothing physical to measure it. We cannot feel it like pressure or temperature.
However, the entropy can be considered as a measure of the disorder of the
molecular distribution of the system. When heat is added to a system, thermal
agitation of molecules increases and the entropy of the system increases. Thus,
the increase of entropy implies a transition from ordered to disordered state of the
system.

5.15 THERMODYNAMIC DEFINITION OF TEMPERATURE


While dealing with Carnot’s cycle, we had attempted to give absolute or Kelvin
scale of temperature in terms of efficiencies of Carnot’s engine. However,
the main question remains that what we should understand by temperature in
thermodynamic terms. To do this, we follow a simple approach. We take two
closed systems, each in thermodynamic equilibrium and put them in contact so
that they interact with each other only in terms of energy transfer as heat and then
look for the property which the two systems should have in common when the
energy transfer as heat ceases.
Let us consider a total system C which consists of two subsystems A and B
separated from each other by a rigid wall which is thermally conducting but does
not allow mass transfer across it. The total system C is an isolated system. The
two subsystems A and B though individually in thermodynamic equilibrium are
not in equilibrium with each other. The situation is depicted in Fig. 5.17 which
also indicates the direction of heat flow.

Rigid wall

Isolated system C
A B
dQ

Fig. 5.17 An isolated system consisting of two closed systems separated from each other
by a thermally conducting rigid wall
The Second Law and Third Law of Thermodynamics 5.29

For the system C, we can write the entropy of the total system in terms of
entropies of subsystems A and B i.e.,
SC = SA + SB (1)
Since entropy is a state function
\ S = S (U, V)
and we can write Eqn. (1) as
SC = SA (UA, VA) + SB (UB, VB) (2)
Since the total system C is isolated, its entropy cannot decrease and since, the
wall between the subsystems A and B is rigid, interaction between A and B cannot
be in terms of energy transfer as work; it can only be in terms of energy transfer
as heat. As a result, the total internal energy
UC = UA + UB = constant (3)
which although remains constant for the isolated system C is shared by the two
subsystems A and B. If r represents the fraction of the total internal energy shared
by A, we can write
UA = rUC;
UB = (1 – r) UC (4)
Since the total system C is an isolated system and the wall separating the
subsystems A and B is a rigid wall, therefore, UC, VA and VB would remain
constant. The only parameter which may vary during the course of interaction
between A and B is only r. When the systems A and B attain thermal equilibrium
with each other, the value of SC should attain its maximum. For this
dSC
____ =0 (5)
dr
From Eqns. (2) and (4), we have
dSC
____
dr ( )
∂SA
= ____
∂UA
dUA
____
VA dr
( )
∂SB
+ ____
∂UB
dUB
____
VB dr

( )
∂SA
= ____
∂UA VA
( )
∂SB
UC + ____
∂UB VB
(– UC) (6)

From Eqns. (5) and (6), it follows that at thermal equilibrium between A
and B

( ) ( )
∂SA
____
∂UA VA
∂SB
= ____
∂UB VB
(7)
5.30 Heat and Thermodynamics

Thus (∂S/∂U)V is the property which the two systems A and B should have in
common when thermal equilibrium is attained. From dimensional analysis, ∂S/∂U
has the dimensions of temperature–1. (S = energy/temperature and U = energy).
Therefore,
1
T ∫ ________ (8)
(∂S/∂U)V
This equation gives us thermodynamic definition of temperature.
We can also establish the direction of heat flow using Eqn. (6) from which it
follows that
1
TA TB (
1
dSC = UC dr ___ – ___ ) (9)

Since TB π TA initially and dSC > 0 for the total isolated system, it follows that
dr > 0 and 1/TA > 1/TB (i.e., TB > TA). UA increases and UB decreases, when TB is
greater than TA i.e., the heat flows from B to A.
Let us take this analysis a little further to learn about the relationship between
dSC
T and U. Since ____ = 0 represents the condition of a maximum, the second
dr
derivative i.e., d SC/dr2 should be negative. Hence from Eqn. (6), we can write
2

d2SC
____
dr2
=
∂2SA
∂UA2( )
_____
VA
( )
∂2SB
UC2 + _____2
∂UB VB
UC2 < 0 (10)

If A and B are identical systems, we get from Eqn. (10)

( )
∂2S
____
∂U2 V
<0 for A and B

or ( )
∂T
___
∂U V
>0 (11)

i.e., the temperature must be a monotonically increasing function of internal


energy at a fixed volume.

5.16 T‐S DIAGRAM


For the heat transferred in a reversible process, we have

dQ(rev) = T dS (1)
f
Q(rev) i, f = Ú T dS (2)
i
The Second Law and Third Law of Thermodynamics 5.31

The integral represents the amount of Area = TdS . dQ


heat Q(rev) added to the system and can
B
be interpreted graphically as the area
under a curve on a T-S diagram, in which A
T
T is plotted along the y-axis and S along
the x-axis, Fig. 5.18.
An isothermal process (dT = 0)
appears as a horizontal line on a T-S
diagram. For a reversible adiabatic SA SB S
process, we have Fig. 5.18 Area under a curve in a T-S diagram
dQ(rev) = 0 represents heat absorbed

dQ(rev)
dS = _____ = 0
T (adiabatic process)
S = constant (3)

Therefore, entropy remains constant during


a reversible adiabatic process and it is called
an isentropic process. Such a process will be T1
T
represented by a vertical line on a T-S diagram.
A Carnot cycle consists of two isothermal T2
and two adiabatic processes. Therefore, on a T-S
diagram it will look like a rectangle, Fig. 5.19. S

Fig. 5.19 Carnot cycle on a T-S


5.17 ENTROPY CHANGE IN AN diagram
IRREVERSIBLE PROCESS
For any process undergone by a system
dQ
___ £ ds (1)
T
dQ
or ds ≥ ___ (2)
T
This is further clarified if we consider the cycle as shown in Fig. 5.20, where A
and B are reversible processes and C is an irreversible process. For the reversible
cycle consisting of A and B,
2 1
dQ dQ dQ
Ú ___ = Ú ___ + Ú ___ = 0
R T 1A T 2B T

2 1
dQ dQ
or Ú ___
T
= – Ú ___
T
(3)
1A 2B
5.32 Heat and Thermodynamics

A Rev.
2
T

Rev.B C
1

Irrev.

S
Fig. 5.20 Pertaining to entropy change in an irreversible process

For the irreversible cycle consisting of A and C, by the inequality of Clausius


2 1
dQ
___ dQ dQ
= Ú ___ + Ú ___ < 0 (4)
T A1 T C 2 T
From Eqns. (3) and (4)
1 1
dQ dQ
– Ú ___ + Ú ___ < 0
B2
T C2 T
1 1
dQ dQ
or Ú ___
T
> Ú ___
T
(5)
B2 C2
Since the path B is reversible
1 1
dQ
\ Ú ___
T
= Ú dS (6)
B2 B2

Since entropy is a state property, so entropy changes for the paths B and C
would be the same i.e.,
1 1
Ú dS = Ú dS (7)
B2 C2

From Eqns. (5) and (7)


1 1
dQ
Ú dS > Ú ___
T
C2 C2
\ For any irreversible process
dQ
dS > ___
T
whereas for a reversible process
The Second Law and Third Law of Thermodynamics 5.33

dQ(rev)
dS = ______
T
For the general case
dQ
dS ≥ ___
T
2
dQ
or S2 – S1 ≥ Ú ___ (8)
1 T

5.18 ENTROPY CHANGE OF A PERFECT GAS IN A


REVERSIBLE PROCESS
Consider 1 mole of a perfect gas at temperature T, pressure P and volume V. If
dW is the work done by the gas in change of state from i to f, dU is the increase
in internal energy and dQ is the heat absorbed by the system, then from the first
law of thermodynamics
dQ = dU + dW
But if CV is the molar specific heat at constant volume and the temperature
of the gas increases by dT, volume increases by dV in the change of state, then
increase in internal energy of gas dU = CV dT and the work done by the gas
dW = PdV
\ dQ = CV dT + PdV (1)
Hence, change in entropy
f f
dQ 1
Sf – Si = Ú ___ = Ú __ (CV dT + PdV)
i T i T
f f
dT P
= Ú CV ___ + Ú __ dV
i T iT
f f
dT R
= Ú CV ___ + Ú __ dV
i T i V

or
Tf
() Vf
(Sf – Si) = CV loge __ + R loge __
Ti Vi() (2)

In Terms of P, T
From perfect gas equation PV = RT, we have
V T Pi
__f = __f × __ (3)
Vi Ti Pf
5.34 Heat and Thermodynamics

Hence, from Eqn. (1)


Tf
Ti () ( )
Tf Pi
Sf – Si = CV loge __ + R loge __ × __
Ti Pf
Tf
Ti () ()
Pi
= (CV + R) loge __ + R loge __
Pf

or
Tf
() Pf
(Sf – Si) = CP loge __ – R loge __
Ti Pi() (4)

In Terms of P, V
From perfect gas equation PV = RT, we have
T P i Vf
__f = __ × __ (5)
Ti Pf Vi
Hence, from Eqn. (1), we get
Pf Vf
Pi Vi( ) ()
Vf
Sf – Si = CV loge ____ + R loge __
Vi

or
Pf
() ()
Vf
(Sf – Si) = CV loge __ + CP loge __
Pi Vi
(6)

5.19 REAL HEAT ENGINES


Reversible Carnot engine with perfect gas (or air) as working substance is not a
practical engine because isothermal processes have to be performed very slowly
and the source and sink are required to have very large heat capacity for heat so
as not to change in temperature during the process. Moreover, the piston should
move without friction. There should also be no loss of heat by conducing from
the system (gas) to surroundings. All
T1
this makes the air-standard Carnot T2
engine very heavy, slow and difficult
to construct. It is only an ideal engine
Q1
with no practical value. p1 A B
In spite of the above difficulties, T1
one may think of constructing a p
reversible Carnot steam engine p2 C C¢ T2
D¢ D
operated in the liquid-vapour region Q2
as shown in Fig. 5.21.
V
The heat Q1 is absorbed from
Fig. 5.21 Carnot steam cycle
a boiler (source) and Q2 rejected
The Second Law and Third Law of Thermodynamics 5.35

to a condenser (sink) at constant temperatures and pressures T1, p1 and T2, p2,
respectively. An efficient turbine can nearly achieve the reversible adiabatic
expansion, since steam expands rapidly and with negligible friction. However,
it is not possible to achieve the reversible adiabatic compression of a two-phase
system.
If the compression is performed slowly the conduction losses occur (dQ π 0).
If it is done rapidly, the temperature of the liquid phase is raised slightly while
the vapour phase gets superheated. Therefore, it is not possible to maintain
equilibrium between the two phases, that is, reversibility cannot be achieved.
To overcome the last difficulty the steam cycle has been modified into what
is known as the Rankine cycle. Other possible practical cycles belong to internal
combustion engines:
(a) Petrol Otto cycle (heat absorbed at constant volume), and
(b) Diesel cycle (heat absorbed at constant pressure).
We shall now discuss their idealized versions where effects of friction,
conduction, etc. are neglected.

5.19.1 Rankine Cycle


The basis of ideal steam engine (Fig. 5.22(a)) is provided by the reversible
Rankine cycle (Fig. 5.22(b)). The cycle is:
(i) a Æ b Starting at point a representing the state of saturated liquid water at
the temperature and pressure T1, p1 of boiler, reversible isobaric isothermal
vaporization of water into saturated steam (point b).
(ii) b Æ c Reversible adiabatic expansion of steam to T2, p2 (point c). This
process corresponds to the passage of steam through the steam engine or
turbine.
(iii) c Æ d Reversible isobaric isothermal condensation of steam into saturated
water (point d).
(iv) d Æ e Reversible adiabatic compression of water to the pressure of the
boiler p1 (point e, only a small change of temperature occurs). This process
is performed by the pump shown in Fig. 5.22(a).
(v) e Æ a Reversible isobaric heating of water to the boiler temperature T1.
This heating takes place after the liquid has been pumped into the boiler
(Fig. 5.22(a)). However, for the cycle to be reversible this heat must be
supplied by a series of heat reservoirs.
The Rankine cycle differs from the Carnot cycle in the step e Æ a. The mean
temperature for this reversible heating being less than the source temperature
T1, the efficiency of the Rankine cycle will be lower than that of a Carnot cycle,
which absorbs all heat Q1 at the single temperature T1 and rejects Q2 at T2.
5.36 Heat and Thermodynamics

p T2
Cylinder T1

Con
Boiler
Denser e a b
T1,p1
Valve Valve T2,p2 p1
T1

Feed
pump d
p2 c
T2

V
Valve Valve
(a) (b)

Fig. 5.22 (a) Theoretical steam engine; (b) Rankine cycle

We can easily compare the efficiencies of Carnot cycle and Rankine cycle by
drawing the T-s diagrams. From Fig. 5.23(a),
Work engine
hCarnot = ____________
Heat absorbed

Area ABC¢D¢
= ___________
Area ABEF
(T1 – T2) (s1 – s2)
= ______________
T1 (s1 – s2)

T 1 – T2
= ______ (1)
T1

T T

T1 A a
B T1 b
F
e
T2 T2
D C d c

F E
s s
(a) (b)

Fig. 5.23 (a) Carnot steam cycle in a T-s diagram; (b) Rankine cycle in a T-s diagram
The Second Law and Third Law of Thermodynamics 5.37

For the Rankine cycle (Fig. 5.23(b)), we can define Tm1 = Q1/Ds1 and
Tm2 = Q2/Ds2 as the mean effective temperatures at which heat is absorbed or
rejected, respectively. Then
Tm1 (sb – se) – Tm2 (sc – sd)
hRankine = ______________________
Tm1 (sb – se)
Tm1 – Tm2
= ________ (2)
Tm1
since de and bc are isentropic processes, sb – sc = sc – sd. Clearly, T2 = Tm2 but
Tm1 < T1. Therefore, we have
h (Rankine) < h (Carnot steam)

5.19.2 Otto Cycle


It is an internal combustion engine, that is, heat is generated inside the cylinder
itself by spark-ignition of petrol vapour and air mixture. The Otto engine is shown
in Fig. 5.24. The engine consists of a cylinder-piston arrangement with two valves.
The cycle (Fig. 5.25) consists of six parts. Four of these involve motion of the
piston and are called strokes.

Intake Exhaust Exhaust


valve valve

Intake Compression Power Exhaust

Fig. 5.24 Otto engine

(i) Intake stroke x Æ a A mixture of petrol (2%) and air (98%) is drawn into the
cylinder at atmospheric pressure through the intake valve by the outward
stroke of the piston.
(ii) Compression stroke a Æ b Both the valves are closed and the piston moves
up the cylinder compressing the mixture rapidly. The compression is
nearly adiabatic and temperature rises considerably.
5.38 Heat and Thermodynamics

(iii) Explosion b Æ c At b the fuel is caused to explode by spark-ignition. The


piston is hardly able to move during the explosion so that volume remains
constant, but a very high temperature and pressure are reached.
(iv) Power stroke c Æ d The adiabatic expansion c-d sets in and the piston is
driven out. There is considerable drop in pressure and temperature.
(v) Valve exhaust d Æ a At the end of the power stroke the exhaust valve opens
and the combustion products flow out rapidly into the atmosphere. The
pressure drops at once to atmospheric.
(vi) Exhaust stroke a Æ x The piston moves up the cylinder throwing out the
remaining gases. The exhaust valve then closes and the intake valve opens
for the start of the same cycle.
The petrol engine (Otto) cycle is
clearly irreversible. Figure 5.25 shows
the idealized air standard Otto cycle
with the working substance as air (petrol
vapour merely helps in ignition) which
behaves like a perfect gas. All processes
are assumed to be reversible (quasistatic
and frictionless). The various steps of
the cycle for one mole of air are then
represented as follows:
(i) x Æ a A quasistatic isobaric Fig. 5.25 Otto cycle
intake stroke of air at pa to a
specific volume v1.
(ii) a Æ b A quasistatic, adiabatic compression stroke from v1 to v2 during
which the temperature rises from Ta to Tb. We have

Ta vg1 – 1 = Tb vg2 – 1 (1)

where g is the ratio of the specific heats.


(iii) b Æ c A quasistatic isomeric rise of temperature and pressure brought
about by the absorption of heat q1 from a series of reservoirs between Tb
and Tc. We have

q1 = cv (Tc – Tb) (2)

(iv) c Æ d A quasistatic, adiabatic expansion (power stroke) producing a drop


in temperature such that

Tc vg2 – 1 = Td vg1 – 1 (3)


The Second Law and Third Law of Thermodynamics 5.39

(v) d Æ a A quasistatic isomeric drop of temperature to Ta and of pressure to


pa bought about by rejection of heat q2 with a series of sinks b between Td
and Ta. We have
Q2 = cv (Td – Ta) (4)
(vi) a Æ xA quasistatic exhaust stroke.
From Eqns. (1) to (4), we can write
Td Ta Td – Ta
__ = __ = v2
( )
g–1
= ______
T T v1 T –T
c b c b

q1 – q2 Td – Ta
h = ______ ______
q1 = 1 – Tc – Tb
v2
( )
g–1
=1– v
1

1
= 1 – ____
g–1 (5)
r
where r = v1/v2 is called the compression (or expansion) ratio.
Note that h (Otto) < h (Carnot) because when r is kept same for both, it is
found that the highest temperature of the Otto cycle exceeds and/or the lowest
temperatures of the Otto cycle is lower than the corresponding source and sink
temperatures for the Carnot cycle.
__ engine r @ 9 and if we put g = 1.5 (for air g = 1.4), we find h (Otto)
In an actual
= 1 – ( 1/÷9 ) = 0.67 = 67 per cent under ideal conditions. In practice efficiency
probably only reaches half this value.
Example 5.1 A cyclic heat engine operates
between a source temperature of 800°C and a sink
temperature of 30°C. What is the least rate of heat
rejection per kW net output of heat engine?
T2
Solution hmax = hrev = 1 – __
T1
30 + 273
= 1 – _________
800 + 273
= 1 – 0.282 = 0.718
Wnet
____
Now, = hmax = 0.718
Q1 Fig. 5.26
1
\ Q1 = _____ = 1.392 KW
0.718
5.40 Heat and Thermodynamics

Q2 = Q1 – Wnet = 1.392 – 1
= 0.392 kW
This is the least rate of heat rejection.
Example 5.2 Which is the more efficient way to increase the efficiency of a
Carnot engine: to increase T1, keeping T2 constant; or to decrease T2, keeping T1
constant?
Solution The efficiency of a Carnot engine is given by
T2
h = 1 – __
T1
If T2 is constant

( )
∂h
___
∂T1 T2
T2
= ___2
T1
∂h
As T1 increases, h increases and the slope ___
∂T1 ( ) T2
decreases (Fig. 5.27a).

Fig. 5.27

If T1 is constant

( )
∂h
___
∂T2 T1
1
= – __
T 1

∂h
As T2 decreases, h increases, but the slope ___
∂T2 ( ) T1
remains constant.

Also,

( )
∂h
___
∂T1 T2
T2
= ___2
T1
The Second Law and Third Law of Thermodynamics 5.41

and
( )
∂h
___
∂T2 T1
T1
= – ___2
T1
Since T1 > T2, therefore

( ) ( )
∂h
___
∂T2 T1
∂h
> ___
∂T1 T2

So, the more effective way to increase the efficiency is to decrease T2.

Perpetual Motion Machine of Second Kind (PMM2)


Q2 Wnet
The efficiency of a heat engine h = 1 – ___ = ____. If Q2 = 0 (Wnet = Q1, or h = 1.00),
Q1 Q1
the heat engine will produce net work in a complete cycle by exchanging heat
with only one reservoir, thus violating the Kelvin-Planck statement of the 2nd law.
Such a heat engine is called a perpetual motion machine of the second kind. Such
a PMM2 is impossible.

5.20 PRINCIPLE OF DEGRADATION OF ENERGY


In universe, due to occurrence of natural processes that are irreversible, entropy
is increasing and ultimately, it may attain the state of maximum entropy when all
the temperatures will be equalized and no work, then, would be possible. It can
be shown that, when entropy is increasing, available energy for doing work, will
be decreasing. Let us consider the example of conduction.
Conduction is irreversible process. Let a quantity dQ of heat be conducted
from the body at a higher temperature T1 to a body at lower temperature T2.
Suppose T0 is the lowest available temperature in the system. Then, the available

( ) T0
( )
energy to start with is dQ 1 – __ . After the transfer, it becomes dQ 1 – __ .
T1
T0
T2
Hence, the loss of available energy in the process of conduction is

( ) ( ) ( )
T0 T0 T0 T0
dQ 1 – __ – dQ 1 – __ = dQ __ – __
T1 T2 T2 T1

( ) dQ dQ
= ___ – ___ T0
T2 T1
Since T1 < T2 and right hand side is positive, so there is a net loss of available
energy. Further,
dQ
___ = (S)T
T2 2
5.42 Heat and Thermodynamics

dQ
___
and = (S)T
T1 1

so that
(S)T > (S)T
2 1

i.e., entropy is increasing while the available energy is decreasing. Since irreversible
processes are continually going on in nature, the available energy of the universe
is continually decreasing. The principle of increase of entropy is equivalent to the
principle of decrease of available energy or degradation of energy. There is yet no
violation of conservation of energy or of first law of thermodynamics. The reason
being that the energy which becomes unavailable is not lost but is transformed to
unavailable form. This principle can be stated now as:
“Whenever an irreversible process takes place, certain amount of energy which
could have been utilized for doing useful work changes to a form in which it
becomes unavailable”.

5.21 THEOREM OF MAXIMUM WORK


For a thermally isolated system consisting of several bodies not in thermal
equilibrium with one another, when equilibrium is established, the system may
do work on some external objects. The transition may, however, occur in different
ways and the final equilibrium state will also be different i.e., energy and entropy
of the final equilibrium states will be different. The total work done on external
objects will depend on the manner in which equilibrium is established. Thus,
there arises a question:
How the final equilibrium state must be reached in order that the system does
the maximum possible amount of work?
Now, let the original energy of the system be U0 and the energy in the
equilibrium state, as a function of entropy in the equilibrium state be U(S). As the
system is thermally isolated, the work done by the system would be
|W| = U0 – U(S)
Here, we have written |W| (instead of W) because W < 0 according to the
convention, as the work is done by the system. Differentiating |W| with respect to
entropy S of the final state (at constant volume),
∂|W|
____
∂S ( )
∂U
= – ___
∂S V
=–T

( ∂U
because dU = TdS – pdV and U = U (S, V) fi dU = ___
∂S ( ) V
( )
∂U
dS + ___
∂V S
dV

∂U
∂S( )
consequently ___
V
( )
∂U
= T and ___
∂V S
=–p )
The Second Law and Third Law of Thermodynamics 5.43

Here T is the temperature of the final state.


Thus, the derivative ∂|W|/∂S is negative i.e., |W| decreases with increasing S. As
the system is thermally isolated, its entropy cannot decrease, the greater possibility
is that W would be greatest when S is constant throughout the process. Hence, we
conclude that the system does maximum work when its entropy remains constant
i.e., when the process of reaching equilibrium is reversible. This is called the
theorem of maximum work.

5.22 NEGATIVE TEMPERATURES


It can be seen that the Boltzmann equation
n
__
( )
E
___
n0 = exp – kT (1)

formally permits the temperature to take on not only positive but also negative
values.
From Eqn. (1), we have
n E
ln __ ___
n0 = – kT
Hence,
E
T = – ______
n
k ln __
n0
It can be seen from this expression that if n < n0 then T > 0.
If, however, an atomic system happened to exist in which n could be greater
than n0, then T will be negative. The more, the energy supplied to the system, the
greater will be the number of particles at the highest energy levels. In the limit,
we can imagine a state in which all the particles will gather at the highest levels.
Such a state is obviously a highly ordered one. We can, therefore, designate
the temperature at which the ordered state sets in by (– 0), in contrast to the
conventional absolute zero (+ 0). The difference between these two ‘zeros’ is that
we approach the first from negative side and the
second from positive side. Thus, the temperatures
of a system are not limited to 0 to • but extend
from + 0 to + •, – • to – 0, + • to – •. The energy
S
dependence of entropy is shown in Fig. 5.28.
T>0 T<0
Consequently, with a paradoxical situation,
to arrive at negative temperatures, instead of
cooling a system below absolute zero (which T = 0 T=±• T=–0
Fig. 5.28
5.44 Heat and Thermodynamics

is impossible), we had to increase its energy. A negative temperature (T < 0)


was found to be higher than an infinitely high temperature. It must be noted that
unstable and non-equilibrium states include not only that corresponding to a
temperature of (– 0), but also all states with negative temperatures. The situations
where n > n0 correspond to all of them whereas for equilibrium we must have the
reverse relationship of these quantities.

5.23 THERMODYNAMIC POTENTIALS


We know that the state of a fixed mass of a thermodynamic system can be
expressed in terms of its pressure P, volume V, temperature T and entropy S.
These are called the thermodynamic variables. These are related as

dQ = dU + PdV (1st law of thermodynamics)


and for a reversible process (1)
nd
dQ = TdS (2 law of thermodynamics)

In the four thermodynamic variables P, V, T and S, only two variables are


independent and remaining two variables are functions of the first two. These
variables are the extensive properties of the system (their values depend on the mass
of the system). From Eqn. (1), combining both the laws of thermodynamics
TdS = dU + PdV
or dU = TdS – PdV (2)
This eqn. represents internal energy of the system in terms of all the four
thermodynamic variables. But, for the complete knowledge of the behaviour of
the system, we require some other relations (in addition to these variables) which
are known as thermodynamic potentials (or energy functions).
Following are the four main thermodynamic potentials:
(i) Internal or intrinsic energy U;
(ii) Helmholtz free energy F;
(iii) Enthalpy or total heat function H;
(iv) Gibb’s free energy G.

5.23.1 Internal Energy U


The internal energy U of a system is a thermodynamic variable. It is the energy due
to molecular constitution and molecular motion of the system. The total energy
of a system is the sum of internal kinetic energy due to its molecular motion and
internal potential energy due to mutual interaction amongst the molecules. It is an
extensive property of the system. It is a unique function of the state of the system
The Second Law and Third Law of Thermodynamics 5.45

i.e., its value depends only on the initial and final states of the system and does not
depend on the path or the process by which the state of the system is changed. In
other words, U is a perfect differential.
If the state of a system is changed from an initial equilibrium state to a final
equilibrium state through an infinitesimal reversible process, the change in internal
energy is given as
dU = TdS – PdV (from Eqn. (1)) (3)
In an adiabatic change of a gaseous system
TdS = 0
and dU = – PdV
Thus, in an adiabatic process, the work done by the gas is equal to the decrease
in internal energy of the system (or the work done on the gas is equal to the
increase in internal energy of the system).

5.23.2 Helmholtz Free Energy F


This is also called the thermodynamic potential at constant volume, and is defined
by
F = U – TS (4)
The free energy F is also a unique function of the state of the system. It is a
perfect differential. It is also an extensive property of the system. Consider an
infinitesimal reversible isothermal change of a system from an initial equilibrium
state 1 to the final equilibrium state 2, then

F1 = U1 – TS1

F2 = U2 – TS2
(since T is constant)
Then, increase in Helmholtz free energy

F2 – F1 = (U2 – U1) – T (S2 – S1)

or (dF)T = (dU)T – TdS


But from the second law of thermodynamics
(dQ)R = TdS
\ (dF)T = (dU)T – (dQ)R
But from first law,
dQ = dU + dW
5.46 Heat and Thermodynamics

\ (dF)T = – (dW)R
or (dW)R = – (dF)T (5)
i.e., decrease in Helmholtz free energy in a reversible isothermal process is equal
to the work done by the system in that process. In other words, in a reversible
isothermal process, the total external work is obtained from the Helmholtz free
energy of the system. Hence, Helmholtz free energy of a system is the energy
which is available for work in a reversible isothermal process.
Thus, this energy is analogous to the potential energy of a system.
From Eqn. (4), we have
U = F + TS (6)
Thus, the internal energy of a system has two parts: (a) Helmholtz free energy
F (which is available for work in a reversible isothermal process) and (b) latent
(or bound) energy TS which is unavailable for the useful work. As the entropy
increases, the unavailable energy increases, hence the available energy for work
decreases.

5.23.3 Enthalpy or Total Heat Function


It is defined by the equation
H = U + PV (7)
The enthalpy H is also a unique function of the state of the system. It is a
perfect differential and also, an extensive property of the system.
Consider an infinitesimal reversible isobaric change of a system from an initial
equilibrium state to a final equilibrium state 2, then
H1 = U1 + PV1
H2 = U2 + PV2 (at constant P)
Increase in enthalpy
H2 – H1 = (U2 – U1) + P (V2 – V1)
or (dH)P = dU + PdV
But dQ = dU + PdV (from 1st law of thermodynamics)
Therefore,
(dH)P = dQ (8)
i.e., in a reversible isobaric process, the change in enthalpy of the system is equal
to the heat extracted (or heat rejected) by the system.
In an adiabatic throttling process, the total initial enthalpy of gas is equal to the
total final enthalpy. In this process, the gas is passed through an insulated porous
The Second Law and Third Law of Thermodynamics 5.47

plug from a constant high pressure to a constant low pressure. Let Pi, Vi be the
initial pressure and volume of the gas and Pf, Vf, the final pressure and volume.
Net external work done by the gas will be (Pf Vf – Pi Vi) which is obtained from
the internal energy of the gas (\ the gas is insulated from the surroundings). If Ui
and Uf be the initial and final internal energies of the gas, then decrease in internal
energy = Ui – Uf. But decrease in internal energy = net external work done by the
gas

or Ui – Uf = Pf Vf – Pi Vi

or Ui + Pi Vi = Uf + Pf Vf

or Hf = Hi (9)

i.e., enthalpy remains constant in an adiabatic throttling process. Enthalpy also


remains constant in an isothermal change of an ideal gas, because dU = 0 i.e.,
U1 = U2 and from Boyle’s law,

P1 V1 = P2 V2

\ U1 + P1 V1 = U2 + P2 V2

or H1 = H2

in an isothermal change.
The locus of points representing the equilibrium states of the same enthalpy of
a thermodynamic system is called an iso-enthalpic curve.

5.23.4 Gibb’s Free Energy G


It is also called the thermodynamic potential at constant pressure. It is defined by
the equation
G = H – TS
= U +PV – TS = F + PV (10)
where the symbols have their usual meanings. G is also a unique function of the
state of the system. It is a perfect differential and an extensive property of the
system.
Consider an infinitesimal reversible isothermal and isobaric change of a system
from an initial equilibrium state 1 to a final equilibrium state 2. Then
G1 = U1 + PV1 – TS1
G2 = U2 + PV2 – TS2
(at constant P and T).
5.48 Heat and Thermodynamics

\ Increase in Gibb’s free energy, in this case is


G2 – G1 = (U2 – U1) + P (V2 – V1) – T (S2 – S1)
or (dG)T, P = dU + PdV – TdS
nd
But from 2 law
(dQ)R = TdS
st
and from 1 law (dQ)R = dU + (dW)R
\ (dG)T, P = (dQ)R – (dW)R + PdV – TdS
= – (dW)R + PdV (11)
If in the above reversible process, the work is done by the system only in
expansion at constant pressure then
(dW)R = PdV
and (dG)T, P = 0 (12)
Thus, in a reversible change at a constant temperature and pressure, the Gibb’s
free energy of a system remains constant if the work is done by the system only
in expansion at constant pressure.
But if, the process is natural (i.e., irreversible), the entropy increases and Gibb’s
free energy of the system decreases (see Eqn. (10)).

5.24 MAXWELL’S RELATIONS


Consider a homogenous system whose volume V is the only external parameter of
relevance. The starting point of our discussion is the fundamental thermodynamic
relation for a quasi-static infinitesimal process
dQ = TdS = dE + pdV (1)
That this equation gives rise to a wealth of other relations will be seen
subsequently.

1. S and V as Independent Variables


Equation (1) can be written as
dE = TdS – pdV (2)
This shows how E depends on independent variations of the parameters S and
V, i.e.,
E = E (S, V)

fi ( )
∂E
dE = ___
∂S V
( )
∂E
dS + ___
∂V S
dS (3)
The Second Law and Third Law of Thermodynamics 5.49

Since Eqns. (2) and (3) must be equal for all possible values of dS and dV,
it follows that the corresponding coefficients of dS and dV must be the same.
Hence,

( )
∂E
___
∂S V
=T
(4)
( )
∂E
___
∂V S
=–p

The important content of Eqn. (2) is that the combination of parameters on


the r.h.s. is always equal to the exact differential of a quantity which is energy
E in this case. Hence, the parameters T, S, p and V which occur in the r.h.s. of
Eqn. (2) cannot be varied completely arbitrarily; there must exist some combination
between them to guarantee that their combination yields the differential dE. To
obtain this connection, it is only necessary to note that the second derivatives of
E must be independent of the order of differential i.e.,
∂2E _____
_____ ∂2E
=
∂V∂S ∂S∂V

or ( )( ) ( )( )

___
∂V S
∂E
___
∂S V

= ___
∂S V
∂E
___
∂V S

Hence, using Eqn. (4), we get

( ) ( )
∂T
___
∂V S
∂p
= – ___
∂S V
(5)

2. S and p as Independent Variables


We have
pdV = d (pV) – Vdp
We now return to Eqn. (2) and transform this into an expression involving dp
rather than dV. Then
dE = TdS – pdV
= TdS – d (pV) + VdP
or d (E + pV) = TdS + Vdp
We can write this as
dH = TdS + Vdp (6)
where we have used the definition
H ∫ E + pV, the enthalpy (7)
5.50 Heat and Thermodynamics

Considering S and p as independent variables

H = H (S, p)

\
∂H
dH = ___
∂S ( ) p ( )
∂H
dS + ___
∂p S
dp (8)

Comparing Eqns. (6) and (8), we get

( )
∂H
___
∂S p
=T
(9)

( )
∂H
___
∂p S
=V

Again, the important aspect of Eqn. (6) is the fact that the combination of
parameters on the r.h.s. is equal to the exactly differential of a quantity, which has
been designated as H. Then
∂2H _____
_____ ∂2H
=
∂p∂S ∂S∂p


( ) ( )
∂T
___
∂p S
∂V
= ___
∂S p
(10)

(using Eqn. (9).

3. T and V as Independent Variables


We now transform Eqn. (2) into an expression involving dT rather than dS. We
can write

dE = TdS – pdV
= d (TS) – SdT – pdV
or dF = – SdT – pdV (11)
using
F ∫ E – TS (12)
The function F is called “Helmholtz free energy”. Considering T and V as
independent variables
F = F (T, V)

∂F
dF = ___
∂T ( ) V
( )
∂F
dT + ___
∂V T
dV (13)
The Second Law and Third Law of Thermodynamics 5.51

Comparing Eqns. (11) and (13), we get

( )
∂F
___
∂T V
=–S
(14)
( )
∂F
___
∂V T
=–p

The equality
∂2F _____
_____ ∂ 2F
=
∂V∂T ∂T∂V
then gives

( ) ( )
∂S
___
∂V T
∂p
= ___
∂T V
(15)

4. T and p Independent Variables


We finally transform Eqn. (2) into an expression involving dT and dp (rather than
dS and dV). Thus, we can write

dE = TdS – pdV
= d (TS) – SdT – d (pV) + Vdp
or dG = – SdT + Vdp (16)

where we have introduced the definition

G ∫ E – TS + pV (17a)
called the Gibb’s free energy.

or G = H – TS (17b)
(using Eqn. (17))
or G = F + pV (17c)
(using Eqn. (12))
Considering T and p as independent variables.
G = G (T, p)

\
∂G
dG = ___
∂T ( ) p ( )
∂G
dT + ___
∂p T
dp (18)
5.52 Heat and Thermodynamics

Comparing Eqns. (16) and (18), we get

( )
∂G
___
∂T p
=–S
(19)

( )
∂G
___
∂p T
=V

The equality
∂2G _____
_____ ∂2G
=
∂p∂T ∂T∂p
then implies

( ) ( )
∂S
– ___
∂p T
∂V
= ___
∂T p
(20)

Thus, we have derived the important relations in Eqns. (5), (10), (15) and (20)
which are summarized as follows

( ) ( )
∂T
___
∂V S
∂p
= – ___
∂S V
(21a)

( ) ( )
∂T
___
∂p S
∂V
= ___
∂S p
(21b)

( ) ( )
∂S
___
∂V T
∂p
= ___
∂T V
(21c)

( ) ( )
∂S
– ___
∂p T
∂V
= ___
∂T p
(21d)

These are known as Maxwell’s relations.

5.25 HOW TO REMEMBER THE MAXWELL’S RELATIONS


∂(p, V)
∂(T, S) ______
______
1. =
∂(x, y) ∂(x, y)
2. We have Jacobian

| |
( ) ( )
∂X
___ ∂Y
___
∂(X, Y)
______ ∂x y ∂x y
=
∂(x, y)
( ) ( )
∂X
___
∂y x
∂Y
___
∂y x

∂X
= ___
∂x ( )( ) ( )( )
y
∂Y
___
∂y x
∂X
– ___
∂y x
∂Y
___
∂x y
The Second Law and Third Law of Thermodynamics 5.53

3. For x and y, we have four variables T, S, p, V. There are in all six possible
conbinations
(T, S), (T, p), (T, V)
(S, p), (S, V)
(p, V)
4. The pairs (p, V) and (T, S) for (x, y) lead to trivial relations, because
from 1,
∂(T, S)
______ ∂(p, V)
= 1 = ______
∂(p, V) ∂(T, S)
Therefore there remain four combinations
(T, V), (T, p), (S, V) and (S, p)
Thus we have four cases as follows.
5. Because from 1 and 2:

( )( ) ( )( )
∂T
___
∂x y
∂S
___
∂y x
∂T
– ___
∂y x
∂S
___
∂x y

( )( ) ( )( )
∂p
= ___
∂x y
∂V
___
∂y x
∂p
– ___
∂y x
∂V
___
∂x y

I. x = T, y = V
This gives (using (5)).

( ) ( )
∂S
___
∂V T
∂p
= ___
∂T V
(first relation)

( ∂T
∂V
∂V
because T and V are independent variables ___ = 0 = ___
∂T )
II. x = T, y = p
This gives

( ) ( )
∂S
___
∂p T
∂V
= – ___
∂T p
(second relation)

III. x = S, y = V

( ) ( )
∂T
– ___
∂V S
∂p
= ___
∂S V
(using (5))

∂S
___ ∂V
= 0 = ___
∂V ∂S

or ( ) ( )
∂T
___
∂V S
∂p
= – ___
∂S V
(third relation)
5.54 Heat and Thermodynamics

IV. x = S, y = p
This gives (using (5))

( ) ( )
∂T
___
∂p S
∂V
= – ___
∂S p
(fourth relation)

The four characteristic functions (U, H, F and V A(F) T


G) and associated Maxwell relations need to be
remembered. A useful mnemonic device for this
purpose (Fig. 5.29), is called as a VAT – VUS
diagram because of the labels on the top and left U G
side. A characteristic function is indicated at the
midpoint of each side and its thermodynamic
coordinates at the ends of the side. So for
example, the Helmholtz function A (i.e., F) is a
function of thermodynamic coordinates V and T, S H P
and the internal energy function U is a function of Fig. 5.29 The VAT-VUS diagram
thermodynamic coordinates V and S.
Now, consider for example, the internal energy function U (V, S). The
differential dU equals the sum of terms including dV and dS. The coefficient of dV
is found by the arrow that connects V to P. Notice that the connection goes against
the arrow, so the coefficient of dV is not P, but – P. Similarly, the coefficient of dS
is found by going in the direction of the arrow that connects S to T (i.e., it is T).
\ dU = TdS – PdV
The Maxwell relations are obtained by applying eqns.
dZ = Mdx + Ndy

( ) ( )
∂M
___
∂y x
∂N
= ___
∂x y

to each of the four thermodynamic potential functions. Consequently, we have


1. dU = TdS – PdV;

hence ( ) ( )
∂T
___
∂V S
∂p
= – ___
∂S V

2. dH = TdS + Vdp;

hence
( ) ( )
∂T
___
∂p S
∂V
= ___
∂S p

3. dA = – SdT – pdV;

hence ( ) ( )
∂S
___
∂V T
∂p
= ___
∂T V
The Second Law and Third Law of Thermodynamics 5.55

4. dG = – SdT + Vdp;

hence
( ) ( )
∂S
___
∂p T
∂V
= – ___
∂T p

(The four equations on the right are Maxwell’s relations).

5.26 THERMODYNAMIC VARIABLES IN TERMS OF


THERMODYNAMIC POTENTIALS
We know that
dU = TdS – pdV (1)
dF = – SdT – pdV (2)
dH = TdS + Vdp (3)
dG = – SdT + Vdp (4)
From Eqn. (1)
∂U
( )
T = ___ ;
∂S V ( )
∂U
p = – ___
∂V S
From Eqn. (2)

( )
∂F
p = – ___ ;
∂V T ( )
∂F
S = – ___
∂T V
From Eqn. (3)
∂H
( )
T = ___ ;
∂S p ( )
∂H
V = ___
∂p S
From Eqn. (4)
∂G
( )
V = ___ ;
∂p T ( )
∂G
S = – ___
∂T p

Thus, the thermodynamic variables S, p, T, V can be obtained from differentiation


of thermodynamic potentials (U, F, G, H) as

( ) ( )
∂F
S = – ___
∂T V
∂G
= – ___
∂T p

( ) ( )
∂U
T = ___
∂S V
∂H
= ___
∂S p

( ) ( )
∂U
p = – ___
∂V S
∂F
= – ___
∂V T

( ) ( )
∂H
V = ___
∂p S
∂G
= ___
∂p T
5.56 Heat and Thermodynamics

5.27 TdS EQUATIONS


We know that the thermodynamic coordinates S, p, T and V of a thermodynamic
system are related to each other and any one of these coordinates can be expressed
as a function of the other two. Thus,
S = S (T, V)
or S = S (T, P)
On this basis, two TdS equations can be obtained.

First TdS Equation


∂P
TdS = CV dT + T ___
∂T ( ) V
dV

Second TdS Equation


∂V
TdS = CP dT – T ___
∂T ( ) P
dP

These can be derives as follows.


Let entropy S of a system be a function of its temperature T and volume V:
S = S (T, V)

\
∂S
dS = ___
∂T ( ) ( )V
∂S
dT + ___
∂V T
dV

or
∂S
TdS = T ___
∂T ( ) ( ) V
∂S
dT + ___
∂V T
dV

But T∂S = ∂Q

and ( )
∂Q
___
∂T V
= CV

\
∂S
TdS = CV dT + T ___
∂V ( ) T
dV

Now, from Maxwell’s relation

( ) ( )
∂S
___
∂V T
∂P
= ___
∂T V

Hence,
∂P
TdS = CV dT + T ___
∂T ( ) V
dV

This is the first TdS equation.


The Second Law and Third Law of Thermodynamics 5.57

Now, let the entropy S of a system be a function of its temperature T and


pressure P.

5.28 EXPRESSIONS FOR C P C V


We know that if the molar specific heats at constant pressure and at constant
volume are CP and CV respectively then, the heat required to raise the temperature
of 1 mole of gas at constant pressure is dQ = CP dT and the heat required at constant
volume is dQ = CP dT and the heat required at constant volume is dQ = CV dT
then, by definition
∂Q
( ) ( )
CP = ___ , CV = ___
∂T P
∂Q
∂T V

and
∂Q
∂T P( ) ( )
CP – CV = ___ – ___
∂Q
∂T V
(1)

But, from second law of thermodynamics


dQ = TdS
Consequently, Eqn. (1) becomes
∂S
CP – CV = T ___
∂T ( ) ( ) P
∂S
– T ___
∂T V
(2)

But entropy S = S (T, V), therefore

( ) ( )
∂S
dS = ___
∂T V
∂S
dT + ___
∂V T
dV

or ( ) ( ) ( )( )
∂S
___
∂T P
∂S
= ___
∂T V
∂S
+ ___
∂V T
∂V
___
∂T P

Multiplying both sides by T, we get

( ) ( ) ( )( )
∂S
T ___
∂T P
∂S
= T ___
∂T V
∂S
+ T ___
∂V T
∂V
___
∂T P

or ( ) ( ) ( )( )
∂S
T ___
∂T P
∂S
– T ___
∂T V
∂S
= T ___
∂V T
∂V
___
∂T P
(3)

From Eqns. (2) and (3)


∂S
CP – CV = T ___
∂V ( ) ( ) T
∂V
– ___
∂T P
(4)

But from Maxwell’s relations

( ) ( )
∂S
___
∂V T
∂P
= ___
∂T V
(5)
5.58 Heat and Thermodynamics

\
∂P
CP – CV = T ___
∂T ( )( )V
∂V
___
∂T P
(6)

The relations (5) and (6) are called the specific heat equations.
Further, from first and second laws,
TdS = dU + PdV

\ ( ) ( )
∂S
T ___
∂V T
∂U
= ___
∂V T
+P

Substituting this value in Eqn. (4), we get

CP – C V = [( ) ] ( )
∂U
___
∂V T
+P
∂V
___
∂T P
(7)

(a) CP – CV For a Perfect Gas


For 1 mole of perfect gas, the equation of state is
PV = RT

\ ( )
∂P
___
∂T V
R
= __
V

and ( )
∂V
___
∂T P
R
= __
P
Hence, from Eqn. (6)
R R
CP – CV = T × __ × __
V P
TR2 TR2
= ____ = ____ = R
PV RT
or CP – CV = R (8)
for a perfect gas.

(b) CP – CV For a van der Waals Gas


For 1 mole of a real (i.e., van der Waals) Gas,

( P + ___Va ) (V – b) = RT
2

where a and b are van der Waals constants.

or
( P + ___Va ) = _____
2
RT
V–b
The Second Law and Third Law of Thermodynamics 5.59

Differentiating the above equation with respect to T at a constant volume

( ) ∂P
___
∂T V
R
= _____
V–b
(9)

Similarly, differentiating with respect to T at a constant pressure


2a ∂V
V
( )
∂T P
RT
(V – b)
∂V
0 – ___3 ___ = – _______2 ___ + _____
∂T P V
R
–b( )
or ( )[
∂V
___
∂T P
RT
_______
(V – b) V2
2a
]R
– ___3 = _____
V –b
R
______
(V – b)
or ( )
∂V
___ = _____________ (10)
∂T P
[
RT
_______
2
(V – b) V
2a
– ___3
]
∂P
Substituting the values of ___
∂T ( ) ( )V
∂V
and ___
∂T P
from Eqns (9) and (10) in Eqn. (6),
we get
R
______
R (V – b)
CP – CV = T ______ × _____________
(V – b) RT
_______
2
(V – b) V
2a
– ___3
[ ]
R
= ________________

( )
2
2a (V – b)
1 – ___3 × _______
V RT
Now, since b << V, therefore (V – b) ª V2, and
2

R
CP – CV = _________
2a
1 – ____
RTV ( )
[ ] 2a
= R 1 – ____
RTV
–1

C – C @ R [ 1 + ____ ]
2a
or P V (11)
RTV
(using binomial theorem)

5.29 CHANGE OF PHASE: EQUILIBRIUM BETWEEN


A LIQUID AND ITS VAPOUR
Substances can exist in three phases solid, liquid and gas. Phase changes involve
consumption of heat called latent heat. The phase-equilibrium line separating the
three phases of a substance appear typically as shown in Fig. 5.30.
5.60 Heat and Thermodynamics

These lines separate solid from liquid,


liquid from gas and solid from gas. The gas
phase is sometimes also called vapour phase.
The three lines meet at one common point
O, called the “triple-point”. At this unique
temperature and pressure, and arbitrary
amounts of all three phases can coexist in
equilibrium with each other. This property
makes the triple point of water a readily
reproducible temperature standard. At point
A, the so called “critical point” the liquid-
gas equilibrium line ends. The volume
changes DV between liquid and gas has then
approached zero: Beyond A there is no further Fig. 5.30
phase transformation, since there exists only
one “fluid-phase”. In going from solid to liquid the entropy (or degree of disorder)
of a substance almost always increases. However, an exceptional case occurs for
solid He3. The corresponding latent heat LS1 is positive and heat gets evolved
in the transformation (solid Æ liquid). In most cases, the solid expands upon
melting so that Vl – VS = DV > 0. But water contracts on melting i.e., DV < 0. The
slope of the melting curve in this case is negative. The three curves OA, OB and
OC are given by the Gibbs free energy equation per unit mass of each phase
gliq – ggas = 0 (Vaporisation)
gsolid – ggas = 0 (Sublimation)
gsolid – gliq = 0 (Fusion or melting)
These three equations must meet at the triple point. The triple point is
independent of pressure. For water it is nearly 0°C. The dotted curve OA¢ is merely
the continuation of OA. It represents the vapour pressure of super cooled liquid.

5.30 CLAUSIUS CLAPEYRON EQUATION


From Maxwell’s relations

( ) ( )
∂S
___
∂V T
∂P
= ___
∂T V
(1)

or ( ) ( )
∂S
T ___
∂V T
∂P
= T ___
∂T V
(2)

But from 2nd law of thermodynamics


TdS = dQ
The Second Law and Third Law of Thermodynamics 5.61

therefore,

( ) ( )
∂Q
___
∂V T
∂P
= T ___
∂T V
(3)

( )
∂Q
Here ___ represents the quantity of heat absorbed or liberated per unit change
∂V T
in volume at constant temperature. This means that at constant temperature, the
heat absorbed or liberated at constant temperature must be the latent heat and
change in volume must be due to a change of state. Considering a unit mass of the
substance, let L be the latent heat when the substance changes in volume from V1
to V2 at constant temperature. Then
dQ = L
and dV = V2 – V1
Substituting these in Eqn. (3), we get

( L
_______
V2 – V1 ) ( )
T
∂P
= T ___
∂T V
(4)

L
_______ dP
or = T ___
V2 – V1 dT
dP _________
___ L
or = (5)
dT T (V2 – V1)
This equation is called the Clausius-Clapeyron latent heat equation.

5.31 FIRST ORDER PHASE TRANSITION AND


CLAUSIUS‐CLAPEYRON EQUATION
We now consider the first order phase transition which are accompanied by the
discontinuities in the first order derivatives of the Gibb’s function. The solid-
liquid and liquid-gas transitions are the first order phase transition.
The thermodynamic state of a homogenous one component system is determined
by two thermodynamic variables for example the pressure P and temperature T.
Let us consider the two phases 1 and 2 in equilibrium, satisfying the condition

G1 (P, T) = G2 (P, T)
(1)
or g (P, T) = G1 (P, T) – G2 (P, T) = 0

Equation (1) can be solved to give the phase-equilibrium curve of P as a


function of T:
(i) The melting curve in the case of solid-liquid transition or
(ii) Pressure curve in the case of liquid-gas transition.
5.62 Heat and Thermodynamics

It is also possible that the equilibrium of three phases is governed by the


condition
G1 (P, T) = G2 (P, T) (2)
G2 (P, T) = G3 (P, T)
These equations have solutions for
specific pair of values (Ptr, Ttr) called
‘triple point’ where three phases are
simultaneously present in equilibrium.
This is depicted in Fig. 5.31 where
regions 1, 2 and 3 are three phases of the
same system. It is to be noted that the
equilibrium of more than three phases
of the same system is impossible.
The molar Gibbs free energies of the
two phases are plotted in Fig. 5.32 as
function of temperature T at constant
pressure. The curves intersect at the Fig. 5.31 Phase diagram of one-component
temperature T0 at which the two phases system
can exist in equilibrium.
At the temperatures below T0, the 1st phase is stable and above T0, the 2nd phase
is stable. This is due to the fact that the stable state is one where G is smaller.
Since entropy is a positive quantity, G must be decreasing function of T. If the
phase 1 is stable for T < T0, then at the point of interaction of the two phases.

( ) ( )
∂G1
____
∂T P
∂G2
> ____
∂T P
(3)

Let us change the temperature of each of the phases by dT and the pressure
by dP in such a way that the two phases continue to be in equilibrium in new
states. Then, in the new condition, the Gibbs free energies of the two phases will
obviously be equal, i.e.,

g (P + dP, T + dT) = 0 (4)


g can be expanded in a Taylor series as
∂g
∂P( )
g (P + dP, T + dT) = g (P, T) + ___
T
( )
∂g
dP + ___
∂T P
dT = 0

The function g (P, T) is always zero at equilibrium, and so

( )
∂g
___
∂P T
( )
∂g
dP + ___
∂T P
dT = 0 (5)
The Second Law and Third Law of Thermodynamics 5.63

Phase 2
G Phase 1

Phase 2
V Phase 1

Phase 2
Phase 1
S

Cp

Phase 1 Phase 2
T
To

Fig. 5.32 The variation of G, V, S and Cp with T during the first order transition

Now, since ( )
∂g
v = ___
∂P T
(6)

and
∂g
s = – ___
∂T P( ) (7)

where v and s are the molar volume and molar entropy respectively, Eqn. (5) gives
the equation of the phase transition boundary

(s2 – s1) dT = (v2 – v1) dP

s2 – s1 Ds
dP ______
___
or = v – v = ___ (8)
dT 2 1 Dv

where s1, v1 and s2, v2 are the molar entropies and molar volumes of the two
phases. Equation (8) is the Clausius-Clapeyron equation.
5.64 Heat and Thermodynamics

The phase transition is accompanied by the evolution or absorption of a certain


quantity of heat (called latent heat). We know that the pressure P and temperature
T remain constant during the phase-transition, therefore,

T (s2 – s1) ∫ (H2 – H1) ∫ L (9)


where L = T (s2 – s1) is the latent heat per unit mass of the transition and the
quantity (H2 – H1) is the difference between the enthalpies of the two coexisting
phases. Substituting this in Eqn. (8), we get
dP __
___ L 1 L
= × _______ = ____ (10)
dT T ( v2 – v1) TDv
Thus, the latent heat L and volume change Dv in a first order transition are
isolated to the slope of the phase equilibrium curve by the Clausius Clapeyron
equation.
During the first order transition, the Gibbs function is continuous but the first
order derivative of the Gibbs function changes discontinuously. These derivatives
are

( )
∂g
(i) the volume v = ___ which changes from v1 to v2 and
∂P T

( )
∂g
(ii) the entropy s = – ___ which changes from s1 to s2.
∂T P
These are demonstrated in Fig. 5.32. There is a jump in both specific volume v
and entropies.
We are also explain the behaviour of other thermodynamic properties e.g.,
specific heat Cp during the first order transition. We know that during the transition
both P and T are constant. Thus, when P is constant, dT = 0 and the specific heat
Cp is

( )
∂s
CP = T ___
∂T P
=•

The variation of CP is also shown in the Fig. 5.32. It remains finite upto
the transition temperature in the phase 1, becoming infinity at the transition
temperature and then returns to finite value for phase 2.
From Eqn. (3), it is clear that s2 > s1 and hence the latent heat is always positive.
The Eqn. (10) may be used for melting, vaporization and sublimation.

5.32 SECOND ORDER PHASE TRANSITION


A phase change of first order is defined as one in which the first order derivatives
of the Gibbs function change discontinuously at the transition. In first order
transition.
The Second Law and Third Law of Thermodynamics 5.65

g2 – g1 = 0

( ) ( )
∂g2
– ___
∂T P
∂g1
+ ___
∂T P
L
= s2 – s1 = __
T

( ) ( )
∂g2
___
∂P T
∂g1
+ ___
∂P T
= v2 – v1

The investigations with liquid Helium by Ehrenfest (1933) have led to the
findings that there are some phenomena where during the phase transition no
latent heat is evolved and there is no change in volume. We then define the second
order phase transition to be one that takes place at constant temperature and
pressure with no change of entropy and volume. The second order phase change
takes place in such a way that there is no discontinuity in the first order derivatives
of the Gibbs function, but the second order derivatives of Gibbs function change
discontinuously.
Now, let us consider a phase change between two phases 1 and 2 in equilibrium
at a pressure P and temperature T, and let
g = G1 – G2
then g (P, T) = 0 (1)
We now change the temperature of each phase by dT and the pressure by dP
such that the two phases continue to be in equilibrium. Then
g (P + dP, T + dT) = 0 (2)
nd
Making a Taylor-series expansion upto the 2 order,

[
∂g
∂P
∂g
g (P, T) + ___ dP + ___ dT
∂T ]
[ ]
2
1 ∂g ∂2g ∂2g
+ __ ____2 (dP)2 + 2 _____ dPdT + ___2 (dT)2
2 ∂P ∂P∂T ∂T
+…=0 (3)
We know that the function g is always zero at equilibrium. If both ∂g/∂P and
∂g/∂T are finite but the sum of the first order terms is zero i.e.,

( ) ( )
∂g
___
∂P
∂g
dP + ___ dT = 0
∂T
(4)

This leads to the Clausius Clapeyron equation and corresponds to the first
∂g ∂g
order phase transition. However, when ___ and ___ are individually zero, we must
∂P ∂T
5.66 Heat and Thermodynamics

∂g ∂g
consider the second order terms. When ___ and ___ are zero both at (P, T) and at
∂P ∂T
(P + dP, T + dT), we have
∂2g
∂g ____
___ ∂2g
= 2 dP + _____ dT = 0 (5)
∂P ∂P ∂T∂P
∂2g
∂g _____
___ ∂2g
= dP + ___2 dT = 0 (6)
∂T ∂T∂P ∂T
Adding Eqns. (5) and (6), we find that the sum of the second order terms of
Eqn. (3) is zero i.e.,
∂2g
____ 2 ∂2g
_____ ∂ 2g
___
2
(dP) + 2 dPdT + (dT)2 = 0 (7)
∂P ∂T∂P ∂T 2

This corresponds to a phase transition of the second order. If such an equality


holds for an arbitrary dP and dT, following conditions must be satisfied:
∂2g
___ >0
∂T2 (8)
∂2g
____ >0
∂P2
and
∂2g ____
___
∂T2 ∂P2
∂2g
( )

∂2g 2
_____
∂T∂P
>0 (9)

∂2G1 _____
_____ ∂2G2
π (10a)
∂T2 ∂T2
∂2G1 _____
_____ ∂2G2
and π (10b)
∂P2 ∂P2
Now, we know that
CP
___
T ( )
∂s
= ___ ;
∂T P (( )∂G
___
∂T P
=–s )
∂2G

∂T [ ( )]
∂G
= ___ – ___
∂T P
= – ____2
∂T
(11a)

isothermal compressibility

( )
k = – __
∂v
1 ___
v ∂P T

or ( )
∂v
kv = – ___ ;
∂P T ( ( )) ∂G
v = ___
∂P T
2
( )
∂ ∂G
= – ___ ___
∂P ∂P T
∂G
= – ____2
∂P
(11b)
The Second Law and Third Law of Thermodynamics 5.67

Volume coefficient of expansion

( )
b = __
∂V
1 ___
v ∂T P

or ( ) [( ) ]
∂V
bv = ___
∂T P

= ___
∂T
∂G
___
∂P T P
2
∂G
= _____ (11c)
∂T∂P
Further, Eqns. (5) and (6), give us two equations. for the phase transition
boundary.
– [ ∂(Dv)/∂T ]P
dP ____________
___ = (12a)
dT [ ∂(Dv)/∂P ]T
and
[ ∂(Ds)/∂T ]P
dP __________
___ = (12b)
dT [ ∂(Dv)/∂T ]P
where Ds = s2 – s1 and Dv = v2 – v1.
With the help of Eqn. (11), Eqn. (12) can be expressed as
v2 b2 – v1 b1
dP __________
___ =
dT v1 k1 – v2 k2
b2 – b1
= ______ (13)
k1 – k2
and
CP – CP
dP _____________
___ 2 1
=
dT T (v2 b2 – v1 b1)

CP – CP
= __________
2 1
(14)
Tv (b2 – b1)
where v1 = v2 ∫ v for the change of the second order. Eqns. (13) and (14) are
known as Ehrenfest equations.
At a second order phase transition, not only the thermodynamic potentials are
continuous but their first derivatives (∂G/∂P)T and (∂G/∂T)P are also continuous,
however, their second derivatives (Eqns. 11) undergo a jump. These are shown
graphically in Fig. 5.33.
Thus, in the first order phase transition, the first derivatives of the Gibbs
function change discontinuously while in the second order phase transition, the
second derivatives of the Gibbs function change discontinuously.
5.68 Heat and Thermodynamics

Fig. 5.33 Variation of G, V, S and CP during the second order transitions

5.33 THE THIRD LAW OF THERMODYNAMICS


We already that an ordered state corresponds to a lower energy of the particles
forming a body. If we cool a body to absolute zero, the maximum conceivable
order would be established in the system and the minimum entropy would
correspond to this state.
On the basis of many experiments conducted at low temperatures, an important
conclusion can be reached:
“at absolute zero of temperature, any changes in state proceed without a change
in entropy”. To remove the lack of uniqueness in the determination of entropy,
Nernst, in 1905, formulated the following rule:
“The entropy of a system at absolute zero is a universal constant which may be
taken to be zero”. This rule is called the third law of thermodynamics. Thus,
entropy of a body at a temperature T is
T
dQ
ST = Ú ___, ST = 0 K = 0 (1)
0 T

Some interesting results follow from the third law.


The Second Law and Third Law of Thermodynamics 5.69

5.33.1 Heat Capacity Vanishes at Absolute Zero


This can be seen by uniting the heat capacity CR (T) associated with a reversible
path R in the form
∂S
CR = T ___
∂T ( ) ( (R)
∂S
= _____
∂ ln T ) (R)
(2)

As T Æ 0, ln T Æ – • and S Æ 0, so that the derivative or CR also tends to zero.


Further, since
T
CR
S = Ú ___ dT (3)
0 T

it follows that heat capacities must approach zero, at least, as rapidly as T for the
integral to coverage and the entropy to become zero.

5.33.2 Absolute Zero is Unattainable by a Finite


Change of Parameters
Consider a system being cooled from Ti to Tf by varying a parameter x from xi to
xf. From second law:
Ti
Cx = xi
ST (xi) – S0 (xi) = Ú _____ dT (4)
i
0 T
(initial state entropy)
and
Tf
Cx = xf
ST (xf) – S0 (xf) = Ú _____ dT (5)
f
0 T
(final state entropy)
Since heat capacities are always positive, we will achieve the greater cooling if
we require the final entropy to be as small as possible. This means that the process
should be adiabatic and reversible (e.g., isentropic demagnetization). Then,
ST = ST
i f

and
Tf Ti
Cx = xf Cx = xi
S0 (xi) – S0 (xf) = Ú _____ dT – Ú _____ dT (6)
0 T 0 T

By the third law


S0 (xi) = S0 (xf) = 0
Therefore,
Ti C Tf
xi Cx f
Ú T dT = Ú ___
___ dT (7)
0 0 T
5.70 Heat and Thermodynamics

If the process continues until Tf = 0, since each of the integrals converges, we


have
Ti
Cxi
Ú ___ dT = 0 (8)
0 T

However, Cxi is greater than zero for Ti π 0, therefore, Eqn. (7) cannot be true
i.e., there is no non-zero solution of the Eqn. (7) and so unattainability of absolute
zero from Eqn. (7), consequently, the third law implies that
“It is impossible to reduce the temperature of a system to absolute zero in any
finite number of operations”.

5.34 APPARENT VIOLATIONS OF THE THIRD LAW


Some systems appear to violate the third law but on closer inspection, the violation
is not real. Glass provides the classic example, Vitreous glass is basically SiO2.
When liquid SiO2 is cooled, it undergoes a phase change at 290 K to become
a crystalline solid. The entropy of this crystalline solid (quartz) decreases
continuously to zero as the absolute zero is approached, in accordance with the
third law.
Molten SiO2 can also be made into glass (amorphous quartz) by rapid cooling.
Figure 5.34 explains the difference between the amorphous and the crystalline
forms of quartz; the former is in a metastable state while the latter is in the
equilibrium state. The third law applies only to systems in equilibrium, and
is, therefore, not applicable to glass. There is, therefore, no question of glass
violating the third law.

Fig. 5.34 Schematic drawing to show that glass is in a metastable state as compared to a crystal
which is in a state of stable equilibrium. Plotted here is the free energy of the system, as a function of
what is called the configuration coordinate. The latter quantity is pretty complicated but we need
not go into all that here. The figure is enough to illustrate that glass is not in an equilibrium state,
and therefore, need not respect the third law. [After G. Venkataraman, “A Hot Story” Universities Press
(India) Ltd. (1993)]
The Second Law and Third Law of Thermodynamics 5.71

Example 5.3 A reversible heat engine is working between temperatures 167°C


and 57°C. if the work done in a cycle is 12 Joules, find how much heat is taken in
at the higher temperature.
Solution Let Q1 be the heat taken in at temperature T1 and Q2 be the heat rejected
at temperature T2, where
T1 = 273 + 167 = 440°K
T2 = 273 + 57 = 330°K
The efficiency of the heat engine is
Q2 T2
h = 1 – ___ = 1 – __
Q1 T1
330 1
= 1 – ____ = __ (i)
440 4
Now, the work done in a cycle is W = 12 Joules which is done by taking the
amount of heat Q = Q1 – Q2.
Thus,
W = JQ
W
or Q = Q1 – Q2 = __
J
12 Joules
= _____________ = 2.87 Calories (ii)
4.18 Joules/Cal
From (i) and (ii)
Q1 – Q2
h = _______
Q1
Q1 – Q2
or Q1 = _______
h
= 2.87 × 4
= 11.48 Cal.
Example 5.4 A Carnot engine working between 100°C and 300°C absorbs 750 J
heat from the high temperature source. Calculate the work done by the engine,
heat rejected and efficiency of the engine.
Solution We have efficiency

W T2 – T1
h = ___ = ______
Q2 T2
573 – 373 200
= _________ = ____ = 0.349
573 573
5.72 Heat and Thermodynamics

Work done by the engine


W = 0.349 Q2
= 0.349 × 750 J = 262 J
Let heat rejected by the engine be Q1. Then
Q2 – Q1 T2 – T1
_______ = ______
Q2 T2
Q1 T1
or ___ = __
Q2 T2
T1 Q2
or Q1 = _____
T2
Thus, heat rejected by the engine
373
= ____ × 750 J
573
= 488 J
Example 5.5 Calculate the difference in temperature of water at the top and
bottom of a rainfall of height 420 m (specific heat of water = 103 cal/kg°C) J = 4.2
J/Cal
Solution Let m kg of water falls from a height h metres. The decrease in potential
energy of water = increases in kinetic energy W = mgh joules. This, energy will
convert into heat. If t°C is the rise in temperature of water,, heat developed
Q = mct cal.
From W = JQ, mgh = J × mcT
gh 4.8 × 420
\ t = ___ = ________3
JC 4.2 × 10
= 0.98°C.
Example 5.6 A lead bullet with velocity 224 m/s strikes a target. If entire energy
produced in impact remains with the bullet, calculate the rise in temperature,
specific heat of lead = 0.03 cal/g°C, J = 4.2 Joule/cal.
Solution The kinetic energy of bullet

= 1/2 mv2 joules


= 1/2 mv2/J calories
If it increases the temperature of the bullet by Dt°C, the heat produced
Q = mCDt cal.
The Second Law and Third Law of Thermodynamics 5.73

From 1st law of thermodynamics,


2
1 mv
__ ____ = mCDt
2 J
1 v2
or Dt = __ _____
2 J×C
1 (224)2
= __ × _____________ = 199°C.
2 4.2 (0.03 × 103)
Example 5.7 100 g steam at 100°C is converted into water at the same temperature.
The latent heat of steam is 540 kcal/kg. Calculate the change in entropy.
Solution Given m = 100 g = 0.1 kg
T = 100°C = 373 K
L = 540 kcal/kg
Decrease in entropy
dQ Q mL
DS = Ú ___ = __ = ___
T T T
0.1 × 540
= ________
373
= 0.145 kcal/K.
Example 5.8 Two bodies of unit mass and specific heat C are at temperatures T1
and T2 respectively (T1 > T2). If they are used as a source of heat and sink for a
reversible engine, show that the maximum work obtained from such arrangement
is
_____
C [ T1 + T2 – 2 ÷T1 T2 ]
Solution Let T be the common temperature of two bodies such that T1 > T > T2.
The change in entropy of the composite system is
T T
mCdT mCdT
DS = Ú ______ + Ú ______
T1 T T2 T
= mC ln (T/T1) + mC ln (T/T2)
= mC [ln (T/T1) + ln (T/T2)]
where m is the mass of the body (here m = 1). Since the engine operates in a
reversible cycle, the entropy change of the composite system is zero. Thus,
C [ln (T/T1) + ln (T/T2)] = 0
or ln (T2/(T1 T2)) = 0
T2
or e° = _____
T1 T2
5.74 Heat and Thermodynamics

T2
_____
or =1
T1 T2
_____
or T = ÷T1 T2
Maximum work is obtained when engine operates as a reversible cycle. The
work done is given by
W = Heat received – Heat rejected
= C (T1 – T) – C (T – T2)
_____
= C [ T1 + T2 – 2 ÷T1 T2 ]

= C [T1 + T2 – 2T]
Example 5.9 Two bodies of equal mass m and specific heat C are at temperatures
T1 and T2 (T1 > T2). They are brought in thermal contact such that they attain an
equilibrium temperature. Show that the change in entropy is

[
(T1 + T2)/2
2mC ln _________
_____
÷T1 T2 ]
Solution Let the equilibrium temperature be T such that T1 > T > T2. From the
law of conservation of energy, we have
mC (T1 – T) = mC (T – T2)
T 1 + T2
or T = ______ (i)
2
The total change in entropy in this case is
T T
mCdT mCdT
DS = Ú ______ + Ú ______
T1 T T2 T

= mC ln (T/T1) + mC ln (T/T2)

= mC ln (T2/(T1 T2)) (ii)


Substituting (i) in (ii)

[
(T1 + T2)2
DS = mC ln ________
4 T1 T2 ]
[ ]
(T1 + T2)/2 2
= mC ln _________
_____
÷T1 T2

= 2mC ln _________
_____
÷T1 T2 [
(T1 + T2)/2
]
The Second Law and Third Law of Thermodynamics 5.75

Example 5.10 A refrigerator operating on the reverse Carnot cycle removes 2000
J/min of heat from a reservoir at 35°C and rejects heat to a reservoir at 100°C.
Calculate the heat rejected to the hot reservoir and the power required.
Solution Efficiency of a Carnot engine
Q2 T2
h = 1 – ___ = 1 – __
Q1 T1
T1
or Q1 = Q2 __
T2
Here T1 = 273 + 100 = 373 K
T2 = 273 + 35 = 308 K
Q2 = 2000 J/min
2000 100
= _____ J/s. = ____ J/s
60 3
So, the heat rejected to the hot reservoir is
T1
Q1 = Q2 __
T2
100 373
= ____ × ____ Joules/s.
3 308
= 40.37 J/s
Hence, the power required
W = Q1 – Q2 = (40.37 – 33.33) J/s
= 7.04 J/s
= 7.04 Watt
7.04
= ____ HP ( 1 HP = 747 Watts)
747
= 0.009 HP
Example 5.11 A Carnot type engine is designed to operate between 480 K and
300 K. Assuming that the engine actually produces 1.2 kJ of mechanical energy
per kilocalories of heat absorbed, compare the actual efficiency with a theoretical
maximum efficiency.
Solution The maximum efficiency
T h – Tc
= ______
Th
480 – 300
= _________ = 37.5%
480
5.76 Heat and Thermodynamics

Energy output
Actual efficiency = ____________
Energy input
1.2
= ________
1 × 4.184
= 28.68%
Example 5.12 A copper can of negligible heat capacity contains 1 kg of water
just above the freezing point. A similar can contain 1 kg of water just below the
boiling point. The two cans are brought into thermal contact. Find the change in
entropy of the system.
Solution For contents of either can
dQ = mCdT
Specific heat of water
C = 1 kcal/kg°K
for water 1 cal = 4.184 J \ 1 kcal = 1 K × 4.184 J
\ mC = 1 kg × 1 K × 4.184 J/kg°K
= 4184 J/K
Since the heat capacities of the two masses are equal, the final temperature will
be the average of initial temperatures
T1i +T2i 273 K + 373 K
T1f – T2f = _______ = _____________ = 323 K
2 2
The change in entropy is
DS = DS1 + DS2
T1f T
dT1 2f dT2
= Ú mC ___ + Ú mC ___
T1i T 1 T2i T2

( )
T1f
T1i
T2i
= mC ln ___ + mC ln ___
T2f ( )
[ (
323 K
273 K ) (
323 K
= (4184 J/K) ln ______ + ln ______
373 K )]
= 100 J/K
Example 5.13 Prove that
∂V
dH = CP dT – T ___
∂T ( ) P
dP + V dP
The Second Law and Third Law of Thermodynamics 5.77

Solution Differentiating S = f (P, T), we get

( )
∂S
dS = ___
∂P T
( )
∂S
dP + ___
∂T P
dT

CP dT
( )
∂S
= ___
∂P T
dP + _____
T
From Maxwell’s relations,

( ) ( )
∂S
___
∂P T
∂V
= – ___
∂T P

Therefore,
CP dT
T
∂V
dS = _____ – ___
∂T ( ) P
dP

Now,
H = U + PV
\ dH = dU + PdV + VdP

= dQ + VdP

= TdS + VdP

[
CP dT ∂V
= T _____ – ___
T ∂T ( ) ] P
dP + VdP

∂V
= CP dT – T ___
∂T ( ) P
dP + VdP

Example 5.14 Deduce an expression for change in entropy of a van der Waals gas
at constant volume and constant temperature.
Solution For a van der Waals gas

( P + ___Va ) (V – b) = RT
2

Let 1 mole of van der Waals gas has pressure P, volume V at temperature T. If
CV is the molar specific heat at constant volume and due to change in state, the
temperature increases by dT and volume increases by dV, then increase in internal
energy of the gas
a
dU = CV dT + ___2 dV
V
(since, internal energy changes in expansion of volume against the internal
pressure a/V2)
5.78 Heat and Thermodynamics

Work done by the gas


dW = PdV
\ Change in entropy
dQ dU + dW
dS = Ú ___ = Ú ________
T T
a


( ___
CV dT + 2 dV + PdV
V
____________________
)
T
a

or
T2
dT (
___
V2 P + 2 dV
V
S2 – S1 = Ú CV ___ + Ú ___________
)
T1 T V1 T
T2 V2
dT RdV
= Ú CV ___ + Ú ______,
T1 T V1 (V – b) ( a RT
P + ___2 = _____
V V –b )
T2
( )
V2 – b
= CV loge __ + R loge ______
T1 V1 – b
At a constant volume,
T2
S2 – S1 = CV loge __
T1
At a constant temperature,

( )
V2 – b
S2 – S1 = R loge ______
V1 – b

Example 5.15 Using the Maxwell’s thermodynamic relations prove that in


reversible isothermal expansion of 1 mole of a van der Waals gas from volume V1
V2 – b
to V2, the amount of heat transferred is RT loge ______ .
V1 – b ( )
Solution Let entropy S of the system is the function of temperature T and
volume V, i.e.,
S = S (T, V)
∂S
dS = ___
∂T ( ) ( )
V
∂S
dT + ___
∂V T
dV

or
∂S
( ) ( )
∂T V
∂S
TdS = T ___ dT + T ___
∂V T
dV

But ( ) ( )
∂S ∂Q
T ___ = ___ = CV
∂T V ∂T V
The Second Law and Third Law of Thermodynamics 5.79

and from Maxwell’s thermodynamic relations

( ) ( )
∂S
___
∂V T
∂P
= ___
∂T V

\
∂P
TdS = CV dT + T ___
∂T ( ) V
dV (i)

Now, for 1 mole of van der Waals gas

( P + ___Va ) (V – b) = RT
2
RT a
or P = _____ – ___2
V–b V
∂P _____
___ R
\ = (ii)
∂T V –b
From Eqns (i) and (ii)
RT
TdS = CV dT + _____ dV
V–b
But for isothermal expansion dT = 0
RT
\ TdS = _____ dV
V–b
V2
RT
\ Ú TdS = Ú ____ dV
V1 V – b

or ( )
V2 – b
Q = RT loge ______
V1 – b

Example 5.16 0.1 kg of steam at 100°C is converted into ice at – 10°C. Calculate
the change in entropy. Given, latent heat of ice = 80 cal/g, latent heat of steam =
540 cal/g, specific heat of ice = 0.5 cal/g°C, specific heat of water = 1 cal/g°C,
log 2.63 = 0.4200, log 2.73 = 0.4363, log 3.73 = 0.5717.
Solution This process is done in four steps: (i) steam at 100°C converts into
water at 100°C (isothermal change) (ii) water at 100°C converts into water at
0°C (iii) water at 0°C converts into ice at 0°C (isothermal change) (iv) ice at 0°C
converts into ice at – 10°C. Given mass of steam m = 0.1 kg = 100 g.
First Step Change in entropy
dQ Q mL1
DS1 = Ú ___ = __ = ____
T T T1

100 × 540
= _________ = 145 cal/K (decrease)
373
5.80 Heat and Thermodynamics

Second Step Change in entropy


Tf
DS2 = mC1 loge __
Ti

( )
273
= 2.3026 × 100 × 1 × log10 ____
373
= 2.3026 × 100 × 1 × [2.4363 – 2.5717]
= – 31.18 cal/K or 31.18 cal/K (decrease)
Third Step Change in entropy
dQ Q mL2
DS3 = Ú ___ = __ = ____
T T T2
100 × 80
= ________ = 29.30 cal/K (decrease)
273
Fourth Step Change in entropy
Tf¢
DS4 = mC2 loge ___
Ti¢

( )
263
= 2.3026 × 100 × 0.5 × log10 ____
273
= 2.3026 × 100 × 0.5 × [2.4200 – 2.4363}
= – 1.88 cal/K or 1.88 cal/K (decrease)
\ Total decrease in entropy
= (145 + 31.18 + 29.30 + 1.88)
= 67.36 cal/K
Example 5.17 1 kg water at 0°C is mixed with 1 kg water at 100°C. Calculate the
change in entropy of the mixture. Given: specific heat of water = 1 kilo cal/kg°C.
log 3.23 = 0.5092, log 2.73 = 0.4362, log 3.73 = 0.5717.
Solution If t°C is the temperature of mixture, then heat given by water at 100°C
= Heat taken by water at 0°C
or 1 × 1 × (100 – t) = 1 × 1 × (t – 0)
\ t = 50°C
Increase in entropy in heating 1 kg water at 0°C (= 273 K) to water at 50°C
(= 323 K)

DS1 = mC loge __
Tf
Ti ()
The Second Law and Third Law of Thermodynamics 5.81

( )
323
= 1 × 1 × 2.3026 log10 ____
273
= 2.3026 × (2.5092 – 2.4362)
= 0.168 kilo.cal/K
The decrease in entropy in cooling 1 kg water at 100°C (= 373 K) to 50°C
(= 323 K) is

DS2 = mC loge __
Tf
Ti()
( )
323
= 1 × 1 × 2.3026 log10 ____
373
= 2.3026 × (2.5092 – 2.5717)
= – 0.144 kilo-cal/K
\ Net increase in entropy of the system
DS = DS1 + DS2 = 0.168 – 0.144
= 0.024 kilo cal/K
Example 5.18 Calculate the change in vapour pressure of water when its boiling
point changes from 100°C to 110°C. Given: Volume of 1 g water vapour =
1640 cm3 and latent heat of vaporization = 540 cal g–1.
Solution dT = 10 K,
T = 100°C = 373 K,
V1 = volume of 1 g of water
= 1 cm3
V2 = volume of 1 g water vapour
= 1640 cm3
\ V2 – V1 = 1639 cm3 = 1639 × 10–6 m3.
L = 540 cal/g = 540 × 4.2
= 2268 J/g
L
\ dP = _________ dT
T (V2 – V1)
2268 × 10
= _______________
373 × 1639 × 10–6
= 0.371 × 105 N/m2
= 0.371 atmospheres
5.82 Heat and Thermodynamics

QUESTIONS
1. Explain the difference between reversible and irreversible processes.
Explain with examples the physical conditions required for a process to be
reversible.
2. What is a real engine? Define efficiency and deduce an expression for it.
3. Discuss the working of Carnot’s ideal engine on P-V diagram and deduce
an expression for its efficiency. Why is it not possible in practice?
4. What is meant by a P-V diagram or indicator diagram? How is the work
done on the system or the work done by the system (during a change of
state) obtained from this graph? Show that the work done on the system or
the work done by the system is the path function.
5. The efficiency of a Carnot engine can be increased in two ways (i) by
decreasing the temperature of the sink T2, keeping the temperature of
source T1 constant, or (ii) by increasing the temperature of source T1,
keeping the temperature of sink T2 constant. Which of these methods is
more effective and why?
6. What is a refrigerator? Show that if an ideal Carnot’s refrigerator working
between the temperatures T1 and T2 (where T2 < T1) extracts heat Q2 at
temperature T2, the work required to run the refrigerator is

(
T1 – T2
W = Q2 ______
T1 )
Obtain an expression for its coefficient of performance.
7. Differentiate between the path function and point function.
8. Define the term entropy. State its units and explain its physical
significance.
9. Show that the entropy is a point function of the state of the system i.e., the
change in entropy of a substance in change of state does not depend on the
reversible path chosen between the two states.
10. Deduce an expression for the change in entropy of a perfect gas in terms
of its pressure, volume and specific heats.
11. State and prove Carnot’s theorem.
12. Explain second law of thermodynamics.
13. Explain the need of second law of thermodynamics. State its both
statements and show their equivalence.
14. Calculate the change in entropy of the universe in a reversible and
irreversible processes.
15. What do you mean by the temperature-entropy diagram? How are the
isothermal and adiabatic curves on this diagram represented? Draw T-S
The Second Law and Third Law of Thermodynamics 5.83

diagram from the P-V diagram of a Carnot cycle and use it to calculate the
efficiency of the cycle.
16. What is thermodynamic scale of temperature? Explain absolute zero on
this scale.
17. Show that the area enclosed by the T-S diagram for a Carnot cycle
represents the available energy.
18. State the principle of increase in entropy and explain it with example.
19. Describe briefly the thermodynamic potentials U, F, H and G. Deduce
Maxwell’s thermodynamic relations from them.
20. Establish thermodynamic variables S, T, P and V in terms of thermodynamic
potentials U, F, H and G.
21. Use Maxwell’s thermodynamic relations to derive the two TdS equations.
Hence find the ratio and difference in two specific heats of a gas.
∂P
22. Show that TdS = CV dT + T ___
∂T ( ) V
dV.

∂V
23. Show that TdS = CP dT – T ___
∂T ( ) P
dP.

∂S
24. Establish the thermodynamic relation ___
∂V ( ) ( )
T
∂P
= ___
∂T V
and use it to deduce

the Clausius-Clapeyron equation


dP _________
___ L
=
dT T (v2 – v1)
Use this expression to explain the effect of pressure on the melting point.
25. Establish Clausius-Clapeyron’s latent heat equation and explain the
effect of increase in pressure on (i) freezing point and (ii) boiling point of
water.
26. Explain the concept of negative temperatures.
27. What is the third law thermodynamics? Explain.
28. Discuss the fact that the Carnot cycle is the most efficient cycle operating
between two fixed temperature reservoirs.
29. What is an internal combustion engine? Explain the working of an Otto
engine with proper diagram and deduce the expression for its efficiency.
What is the limit of its efficiency?
30. At atmospheric pressure (1.01 × 105 N/m2) and 100°C, 1 cm3 of water
forms 1671 cm3 of steam. Latent heat of steam is 540 cal/gm. Calculate the
external work needed to convert 1 g of water into steam at one atmospheric
pressure. What is the increase in internal energy of water? Mechanical
equivalent of heat = 4.2 J/cal. [Ans. 40.16 cal., 499.84 cal.]
5.84 Heat and Thermodynamics

31. Calculate the efficiency of Carnot engine working between 127°C and
27°C. [Ans. 25%]
32. What is the efficiency of a Carnot engine working between 300°C and
15°C? Calculate the work done by the engine if its absorbs 100 cal. if heat
at the high temperature. [Ans. 0.497, 207.9 J]
33. An engine operating between 27°C and 227°C develops 74600 watt power.
Calculate
(i) the efficiency of the engine
(ii) amount of heat absorbed
(iii) amount of heat rejected
[Ans. (1) 40%, (ii) 186500 J/s, (iii) 111900 J/s]
34. Assuming that a domestic refrigerator works as a reversible engine between
the temperatures 0°C and 17°C, find the electrical energy required to freeze
1 kg of water at 0°C (latent heat of ice 80 cal/gm, 1 cal = 4.18 J)
[Ans. 2.08 × 104 J]
35. 20 g water at 70°C is mixed with 10 g water at 40°C. Find the change
in entropy of the system. (Given log 3.13 = 0.4786, log 3.33 = 0.5224,
log 3.43 = 0.5353) [Ans. 0.4145 cal/K (increase)]
36. 1 kg of ice at 0°C is converted into water at the same temperature. What is
the change in entropy? [Ans. 293 cal/K (increase)]
37. 10 g of steam at 100°C is converted into water at the same temperature.
Find the change in entropy. [Ans. 14.48 cal/K (decrease)]
38. Calculate the change in entropy when 1 kg ice at – 10°C is changed into
the steam at 100°C. Given specific heat of ice = 0.5 Kcal/kg°C, latent heat
of ice = 3.4 × 105 J/kg, latent heat steam = 22.68 × 105 J/kg, J = 4.2 Joule/
cal) [Ans. 8.71 × 103 J/K (increase)]
39. 10 g water at 20°C is converted into ice at – 10°C. Calculate the change in
entropy. (log 2.63 = 0.4200, log 2.73 = 0.43622, log 2.93 = 0.4669)
[Ans. 3.824 cal/K (decrease)]
40. Give example(s) of apparent violation of the third law of
thermodynamics.

OBJECTIVE TYPE QUESTIONS


1. The necessary condition for a reversible process is
(a) the loss in energy in complete cycle must be zero.
(b) the process should be done suddenly.
The Second Law and Third Law of Thermodynamics 5.85

(c) the process should be done slowly so that the working substance
remains in thermal and mechanical equilibrium with its surroundings
in each intermediate state.
(d) the loss in energy must be zero and the process must be quasistatic.
2. The reversible process is
(a) the flow of heat from one body to the other when the two bodies at
different temperatures are kept in contact.
(b) the slow expansion of a gas enclosed in an insulated cylinder kept in
contact with a heat source.
(c) the adiabatic compression of a gas.
(d) the slow movement of frictionless piston out of the perfectly insulating
cylinder in contact with a heat source.
3. An ideal gas is compressed isothermally the process is
(a) reversible
(b) irreversible
(c) either reversible or irreversible
(d) nothing can be said till other restrictions ate known
4. The area of P-V diagram of a Carnot cycle represents
(a) rejected energy per cycle (b) absorbed energy per cycle
(c) work done per cycle (d) change in volume per cycle
5. If the temperature of source and sink and T1 and T2 respectively, the
efficiency of Carnot engine is
(a) 1 – T1/T2 (b) (T1 – T2)/T
(c) T1/T2 (d) 1 – T2/T1
6. A Carnot engine works between ice point (0°C) and steam point (100°C) .
Its efficiency is
(a) 36.6% (b) 37.3%
(c) 26.8% (d) 73.2%
7. If the temperature of sink is 0°K, the efficiency of Carnot engine will be
(a) 100% (b) 50%
(c) 10% (d) 25%
8. According to Clausius theorem
dQ dQ
(a) ___ > 0 (b) ___ < 0
T T
dQ dQ
(c) ___ = constant (d) ___ = 0
T T
5.86 Heat and Thermodynamics

9. The correct relation is


dS
(a) dQ = ___ (b) dQ = SdT
T
dV
(c) dQ = dU + dW (d) dQ = dU + ___
P
10. In an irreversible process
(a) the internal energy of universe increases
(b) the temperature of universe increases
(c) the entropy of universe increases
(d) the heat and internal energy of universe increases
11. The entropy change in a complete Carnot cycle is
Q2
(a) zero (b) – ___
T2
Q1 (Q1 – Q2)
(c) + ___ (d) ________
T1 T1
12. The entropy of a system in an reversible adiabatic process
(a) increases (b) decreases
(c) both (a) and (b) (d) none of these
13. The entropy of gas in irreversible adiabatic process
(a) decreases (b) increases
(c) remains unchanged (d) none of these
14. The entropy of a system in reversible cycle
(a) remains unchanged (b) remains zero
(c) increases (d) decreases
15. According to the principle of increase in entropy, in a natural process
(a) DS = 0 (b) DS > 0
(c) DS < 0 (d) DS £ 0
16. Entropy remains constant in
(a) adiabatic process (b) isothermal process
(c) isobaric process (d) cyclic process
17. The second law of thermodynamics is
dQ dQ
(a) ___ ≥ 0 (b) ___ > 0
T T
dQ
___ dQ
___
(c) <0 (d) =0
T T
The Second Law and Third Law of Thermodynamics 5.87

18. On T-S diagram the Carnot cycle is a


(a) square (b) rectangle
(c) parallelogram (d) none of these
19. A temperature difference of 10°C on thermodynamic scale is equal to
(a) 284 K (b) 273 K
(c) 10 K (d) 263 K
20. The thermodynamic scale of temperature is based on
(a) Carnot engine (b) entropy
(c) gas (d) energy
21. Enthalpy is
(a) H = U – TS (b) H = U + V
(c) H = U – PV (d) H = U + PV
22. The following is not a function of state of the system
(a) W (b) V
(c) F (d) H
23. In a reversible process, Gibb’s function of the system
(a) remains same (b) decreases
(c) increases (d) both (a) and (b)
24. Clausius-Clapeyron equation is
dP dP L
(a) ___ = L (V2 – V1) (b) ___ = _________
T dT T (V2 – V1)
L dP T
(c) dT = __ (V2 – V1) dP (d) ___ = _________
T dT L (V2 – V1)
25. The correct relation is
C P ET
(a) CP – CV = ES – ET (b) ___ = ___
CV ES
CP ES
∂V
(c) CP = ___
∂T ( ) P
(d) ___ = ___
CV ET

Answers
1. (d); 2. (d); 3. (a); 4. (c); 5. (d); 6. (c); 7. (a);
8. (d); 9. (c); 10. (c); 11. (a); 12. (a); 13. (c); 14. (a);
15. (b); 16. (a); 17. (d); 18. (b); 19. (c); 20. (a); 21. (d);
22. (a); 23. (a); 24. (b); 25. (d).
6
Chapter

Radiation

6.1 INTRODUCTION
The heat transfer takes place from the region of higher temperature to the region
of lower temperature of a medium.
There are three modes of heat transfer
1. Conduction is the transfer of heat energy by some sort of collision without
gross movement of any part of the body, due to temperature difference
between two parts of a system, or between two systems in direct contact.
2. Convection is the transfer of heat energy due to the motion of heated
fluid particles from the region of higher temperature to that of lower
temperature.
3. Radiation is the transfer of heat energy due to continual emission of energy
from the surface of all bodies in the form of electromagnetic waves.
Usually more than one mode of heat transfer is present at the same time,
although one of the three may be predominant.
Radiation does not require presence of medium for the heat transfer.

6.2 HEAT WAVES


The heat transfer by conduction and convection depends upon the existence of
a medium to convey heat. The third mode of heat transfer, radiation, does not
require any medium. In this process energy, called the radiant heat or thermal
radiation, is continually emitted from the surface of all bodies and is in the form
of electromagnetic waves.
The heat waves occupy the wavelength region extending from about 10–4 cm
to 10 cm in the electromagnetic spectrum shown in Fig. 6.1. These waves travel
in vacuum with the velocity of light. When eclipses of the sun occur, the heat is
cut off at the same instant as the light.
Radiation, like light, travels in straight lines in a homogeneous medium
and follows the ordinary rules of photometry which depends on the rectilinear
propagation (for example, inverse square law). It is reflected and refracted
6.2 Heat and Thermodynamics

Ultraviolet

Infrared
] Visible
x-rays Heat Radio waves
Gamma-rays

22 14 6
10 10 10 vib/sec
–12 –4
10 10 cm
Fig. 6.1 Electromagnetic spectrum

according to the same laws as light. Convex lenses of rock salt or quartz can
converge parallel radiation at their foci. A prism of rock salt or quartz can analyse
a beam of radiation into its various wavelength constituents forming a spectrum
in the invisible region beyond the red. The short wavelength heat radiations
are called infrared radiations. Like light, radiation is absorbed by dark, rough
surfaces, and reflected by light, smooth surfaces. It exhibits the phenomena of
interference, diffraction and polarization. Some of the characteristic properties of
heat radiations are due to the fact that their wavelengths are larger than those of
the visible light.
The radiation from a gas of free molecules contains waves of few frequencies
which are characteristic of the molecules of the gas. On the other hand, the
radiation from a solid or liquid contains waves of all frequencies, with relative
amplitudes (large amplitudes in the region
of heat waves) that depend mainly on the B A
temperature of the emitting body rather than
on the type of molecules involved. We shall
be concerned here only with the radiant
energy emitted by solids and liquids.

6.3 DETECTION AND


MEASUREMENT OF
THERMAL RADIATION
A simple detecting instrument is the
differential air thermometer first used by
Leslie. It consists of two similar glass bulbs
A and B containing air at ordinary pressure.
The U-tube connecting the bulbs is filled with
some nonvolatile liquid like sulphuric acid,
Fig. 6.2. Bulb B is blackened. When heat
Fig. 6.2 Differential air thermometer
Radiation 6.3

radiation falls on B, while A is shielded Hot


from it, the temperature of the air inside
Bi Bi Bi
B increases causing a difference of levels
of the liquid in the manometer.
Melloni introduced thermopile for Sb Sb Sb
detecting and measuring the radiant energy.
It usually consists of a set of antimony- Cold
bismuth thermocouples connected in
series, Fig. 6.3. The hot junctions are
blackened to form the face B which is
G
exposed to the radiation.
In an actual thermopile the couples are Fig. 6.3 Principle of thermopile
made of short bars and a large number of
them are arranged in a rectangular pile,
A
Fig. 6.4. The hot junctions fall on the
face B which is blackened. The rods are
insulated from one another by placing
mica along their lengths. This thermopile
is mounted in a long metal cone, Fig. 6.5,
so that stray radiations do not reach the
black face B. The face A is shielded by a
metal cap C. The thermopile is connected
to a sensitive galvanometer. B
Langley used a bolometer which
in principle is an extremely sensitive Fig. 6.4 The thermopile
platinum resistance thermometer of low
thermal capacity. It consists of a grid
made of very thin platinum strips joined
in series, Fig. 6.6. The grid is blackened
to absorb radiation. Four such grids, G
P, Q, R and S are connected to form a
Wheatstone bridge, Fig. 6.7. Initially the
bridge is balanced so that no current flows
through the galvanometer G. When the
radiation is allowed to fall on the grid P,
its temperature increases. The resistance C A B
of P consequently changes and a current
flows through G, which is a measure of
the radiation falling on the grid P.
In the above arrangement some
radiation passes through the gaps Fig. 6.5 Mounted thermopile
6.4 Heat and Thermodynamics

P Q
G

R S

Fig. 6.6 Bolometer grid Fig. 6.7 Wheatstone bridge

between adjacent strips. To utilize this radiation


the grid S is placed behind P such that the radiation
passing through the gaps in P falls on the strips G

of S, Fig. 6.8. When both P and S get heated up,


their resistances increase and give rise to currents
which flow in the same direction through G. This S P R Q
increases the sensitivity of the arrangement. For
spectrum work, where a line has to be detected, a
single narrow strip of platinum instead of a grid is
used. This is called a linear bolometer.

6.4 EMISSION AND ABSORPTION


OF RADIATION
Fig. 6.8 Bolometer grids in
Experience shows that all substances emit and the bridge circuit
absorb radiant energy at a rate which depends on the
absolute temperature and on the nature of the surface. The electromagnetic waves
in the interior of every body are similar if the bodies are at the same temperature.
That is, the interior of a body is uniformly filled with waves of all frequencies
travelling in all directions, and for a given temperature the wave amplitude of any
particular frequency is the same for every substance. The observed difference in
the amount of radiation emitted by different bodies at the same temperature arises
from the nature of the surface.
When the thermal radiation Q falls upon a body equally from all directions,
the radiation is said to be isotropic. A portion A of the incident radiation may be
absorbed, another portion R, reflected, and the remaining portion T transmitted.
That is
Q=A+R+T (1)
Radiation 6.5

We can write it as

a+r+t=1 (2)

where a = A/Q is the absorptivity, r = R/Q the reflectivity, and t = T/Q the
transmittivity.
For some substances, such as lampblack, the absorptivity is nearly unity.
For theoretical reasons it is useful to think of an ideal substance which absorbs
all radiation falling on it; a = 1, t = 0, r = 0. Such a substance is called a
blackbody. Lampblack, platinum black and an enclosure with a pinhole are good
approximations to a blackbody.
When a = 0, r = 1, t = 0, we have a perfect reflector. Highly polished surfaces
are good approximations to a perfect reflector.
Since the radiant energy absorbed can contribute to increase in the internal
energy of a body, and since the rate of emission of radiation by a body depends on
its temperature, a good absorber is also a good emitter. To see this a Leslie’s cube
(cubical tin box) is taken. It is coated with lampblack on face 1, with black enamel
on face 2, with dull white paint on face 3 and the face 4 is brightly polished.
Boiling water is kept in the cube and radiation is detected with a thermopile at
the same distance from each face in turn. It is found that face 1 radiates strongly,
faces 2 and 3 radiate less strongly, and face 4 radiates feebly. Thus the lampblack
surface is a good absorber and a good emitter while polished surface is a poor
absorber and a poor emitter.
We now describe Ritchie’s experiment (1833) which shows that the emitting
power of a surface is proportional to its absorbing power. A Leslie unit cube Z
is placed between two cubical bulbs, X and Y, of a differential air thermometer,
Fig. 6.9. The faces, facing each other, are [polished (P) or blackened (B) as shown
in the figure. When the water in Z is heated, it is found that there is no change in
the manometer level.

P B P B

X Z Y

Fig. 6.9 Ritchie’s experiment


6.6 Heat and Thermodynamics

This result can be explained as follows. Suppose e(b) is the energy emitted per
second per unit area of surface by the blackened face, and e(p) by the polished
face. The corresponding absorptivities are a(b) = 1 and a(p). The energy incident
on X per second is e(b) of which a fraction a(p) e(b) is absorbed by it. On the other
hand, the energy incident on Y per second is e(p), all of which is absorbed by it.
Since the manometer stays steady.
a(p) e(b) = e(p)
or
e(b) ___
___ 1 a(b)
___
= = (3)
e(p) a(p) a(p)
Thus, the emitting power of a surface is proportional to its absorbing power.

6.5 PREVOST’S THEORY OF EXCHANGES


Consider two bodies A and B. Suppose B is hotter than A. As a result of radiation
between the two bodies, the temperature of B falls while that of A rises. Thus, the
heat is transferred from B to A by radiation. Pierre Prevost of Geneva suggested
in 1792 that the transfer must be considered as a net transfer; that is, each body
radiates energy to and absorbs energy from the other. In our example, B radiates
more energy than it absorbs, while A absorbs more than it radiates.
Prevost said that the amount of energy radiated by a body increases with
temperature, but is not affected by the presence of surrounding bodies. The
change of temperature in a body is due to the exchange of radiant energy with
surrounding bodies.

6.6 EMISSIVE POWER


If the radiant energy is dispersed by a quartz prism, Fig. 6.10, the energy is
found to be distributed continuously among different wavelengths, Fig. 6.11. At
temperatures below about 500°C, most of the energy is associated with infrared
waves.
The rate at which energy is radiated at a wavelength l, per unit surface area, is
called the monochromatic emissive power el. The total emissive power, e, is the
energy radiated per unit time per unit surface area, at all wavelengths. Thus, the
areas under the curves of Fig. 6.11 give e at the corresponding temperatures,
l=•
e= Ú el dl (4)
l=0

Let Ql dl be the radiant energy incident on a body with wavelengths between


l and l + dl and let Al dl be the energy absorbed. Then the absorption coefficient
or the monochromatic absorption power of the surface, al is defined by
Radiation 6.7

Ammeter

White
screen

Infrared
Lens
Thermopile
Visible Red
Ultr Violet
Prism avio
+ let


Carbon arc

Fig. 6.10 Spectrum analysis

Visible

30
l

00
e

20 °
Ultra Infrared 00 K
violet °K
l 1 Energy

60 0 °K
00 100
50 °
00 K 0 2 4 6
°K 4
400 l × 10 cm
e

0 °K
K
3000 °

0 1 2 3
4
l, Wavelength × 10 cm

Fig. 6.11 Distribution of the energy emitted by a hot body at different temperatures

Al dl Al
al = _____ = ___ (5)
Ql dl Ql
Similar definitions for rl and tl can be given.

6.7 RADIATION IN AN ENCLOSURE


All objects are continually absorbing and emitting radiation. For a body to be in
equilibrium with its surroundings, the radiation it is emitting must be equivalent
(in wavelength and energy) to the radiation it is absorbing. A few important results
for the radiation in an enclosure are as follows:
1. A body placed in an enclosure at a constant temperature will ultimately
acquire the temperature of the enclosure (thermal equilibrium).
2. The radiation inside an enclosure is homogenous and isotropic.
6.8 Heat and Thermodynamics

3. Radiation inside an enclosure depends only on its temperature and is


independent of the nature of the walls or of the presence of any body
inside it.
4. Radiation inside the enclosure is identical in every respect with the
radiation emitted by a blackbody.

6.7 KIRCHHOFF’S LAW


Consider a body B, with any type of surface, at equilibrium in an enclosure of
uniform temperature. Suppose the energy dQl in the wavelength range l to l + dl,
which depends only upon the temperature of the enclosure, is falling on unit area
of B per second. We must have for a given wavelength,
energy absorbed by B = energy emitted by B
al dQl = el dl (6)

where al is the absorptive power and el the emissive power of the body. Since
dQl/dl depends onlt on the temperature,
el
__
al = constant (7)

for a given temperature. Thus, the ratio of the emission power to the absorptive
power for radiation of a given wavelength is the same for all bodies at the same
temperature. This is called Kirchhoff’s law (1859).
If the enclosed body B happens to be a perfect blackbody with emissive power
e(b) and absorptive power a(b)
l
= 1, Eqn. (6) becomes
dQl = e(b)
l
dl (8)
Combining (6) and (8), we can write Kirchhoff’s law as
el
Œl ∫ ___ = al (9)
e(b)
l
el
__ (b)
al = el = constant (10)

The ratio el/e(b)


l
is called the emissivity Œl, and as defined here is equal to the
absorptive power. Note that e(b) l
increases with temperature. It remains constant in
the sense that it has the same value for all kinds of surface at a given temperature
and wavelength l.
An ideal surface for which al = constant π 1, that is, for which el is exactly
the same fraction of e(b)
l
at all wavelengths is called a grey-body. If al varies with
wavelength we say it is coloured.
Radiation 6.9

If Eqns. (6) and (8) are integrated from l = 0 to l = •, we get from (4)

e
___ =a=Π(11)
e(b)
Then for Kirchhoff’s law the specification of the wavelength is not needed and
Πis called the total emissivity. The Πdepends on the total emissive power e, or
the area under curves such as shown in Fig. 6.11. However, it is clear from Fig.
6.11 that Eqn. (11) is true only if Πand a have the same wavelength distribution
or if they are independent of wavelength (grey-surface).
We have proved Kirchhoff’s law for bodies in an enclosure. By Prevost’s theory
of exchanges, the radiation emitted by a body depends only on its temperature and
not on its surroundings. Therefore, Kirchhoff’s law applies under all conditions.
If a body shows strong absorption for a given type of radiation strongly, it also
emits a similar type of radiation strongly. Thus, red glass, which looks red as
it absorbs green light strongly, glows with a green light when heated to a high
temperature.
Note that Eqn. (11) is same as Eqn. (3). Therefore, Kirchhoff’s law simply
says that a good absorber is also a good emitter. It is in agreement with Ritchie’s
experiment.
An important application of Kirchhoff’s law is in the field of spectroscopy.
Fraunhofer in 1823 measured the exact wavelengths of the dark lines observed by
him in the spectrum of sunlight. Their origin could be understood only 35 years
later when Kirchhoff in 1859 announced his famous law according to which if a
body absorbs light of a certain wavelength, it also emits this same wavelength.
Thus, the Fraunhofer lines are due to the absorption of certain wavelengths
by sun’s atmosphere. One can find out what elements exist in the sun if one
can observe what elements on earth give bright lines on emission at the same
wavelengths as the dark lines in the solar spectrum. It is found, for example, that
the wavelengths of the emission D-lines of sodium exactly coincide with those
of dark lines of Fraunhofer showing that there is sodium in the solar atmosphere.
Such identifications for various elements have been done for thousands of
observed dark lines.

6.8 BLACKBODY
Kirchhoff defined a blackbody or a full radiator as a body which, at all temperatures,
absorbs all the radiation falling on it; that is

a(b)
l
=1

for all wavelengths.


6.10 Heat and Thermodynamics

From Section 6.7 that inside an enclosure the radiation in any given direction
must be the same as in any other direction (isotropic); it must be the same at
every point (homogenous); and it must be the same in all enclosures at a given
temperature, irrespective of the materials composing them. Moreover, all these
statements hold for each spectral component of the radiation.
A simple relation exists between the radiation in an enclosure and the radiation
emitted by a blackbody. Suppose that such an enclosure contains a black surface.
Then the radiation leaving this surface will consist entirely of radiation emitted
by it because it reflects none of the radiation that falls upon it. Therefore, the
radiation emitted by a black surface or body in any direction is the same as the
radiation travelling in any one direction in an enclosure at the same temperature.
From now on both will be called as blackbody radiation.
A perfect blackbody does not exist in nature,
although lampblack and platinum black are good
approximations. Kirchhoff devised a blackbody with
the help of a hollow enclosure or cavity, Fig. 6.12.
A small hole is cut in a large hollow enclosure, such
as a sphere. The radiation incident upon this hole is
partially absorbed by the inner wall of the enclosure
and the rest is diffused within the enclosure. The
radiant energy, having once entered the enclosure Fig. 6.12 Hollow enclosure
through the hole, continues to be absorbed, reflected blackbody
and reradiated. Only a very small amount of this
diffused radiation finds its way out through the hole; that is al @ 1 for each l, as
required for a blackbody.
The hole of the enclosure behaves very much like
a perfect radiator. The diffused radiation coming
out from such an isothermal enclosure is called the
blackbody (or full) radiation. The term ‘full’ indicates
that the radiation emitted by a non-blackbody may P O
be deficient in some wavelengths and so may not be
complete or full.
Fery has constructed a blackbody, Fig. 6.13, on
this principle. To avoid direct radiation from the
Fig. 6.13 Fery’s blackbody
surface opposite the hole O, which would make the
body not perfectly black, be introduced a conical projection P.

6.9 INTENSITY AND ENERGY DENSITY


Let dA be the small surface element of a radiating body. Describe about dA a
hemisphere of radius r, Fig. 6.14. Consider an area dS on this hemisphere. The
radius to dS makes an angle q with the radius which is normal to dA. The rate dQ/
Radiation 6.11

dt at which radiant energy falls upon dS is proportional to (1) the area dS, (2) the
area dA of the radiator, and (3) 1/r2, due to the inverse-square law. It will also
depend upon q, because the apparent area of dA, as seen from dS, is dA cos q.
Therefore,
dQ
___ dS
= I dA cos q ___2 (1)
dt r
where constant of proportionality I is called the total intensity of radiation from
dA in the direction of dS. We can write
dQ/dt
I = __________ (2)
dA dw cos q
where dw = dS/r2 is the solid angle which the area dS subtends at O. If the radiation
considered is in the wavelength region l to l + dl, we get the monochromatic
intensity of radiation Il.
When I is the same in all directions, there is a simple relation between I and
the total emissive power e. The area of the ring of width rdq on the hemisphere is
2pr sin q rdq. The flux of energy dQ/dt through this ring is
2pr sin q rdq
( )
dQ
___
dt ring
= I dA cos q ____________
r2
(3)

= 2p I dA cos q sin q dq (4)


Since e is the total rate of flux of energy per unit area from a surface, we have
1 dQ
( )
e = ___ ___
dA dt hemisphere

p/2
= 2p I Ú cos q sin q dq (5)
0
or
e = p I ( sin2q )p/2
0
=pI (6)
Similarly,
el = p Il (7)
If a surface obeys the cosine law [dQ/dt μ cos q, Eqn. (1)], so that the intensity
I of the radiation emitted by it is the same in all directions, then a simple relation
exists between e and the energy density Y in the emitted radiation at a point near
the surface dA. In Fig. 6.14 allow the radius r of the sphere to be very large. Then
all rays from points on dA to points on dS may be regarded as parallel. The radiant
energy emitted in 1 second from dA, in the direction of dS, will be contained in a
cylinder whose length is the velocity of light c and cross-sectional area dA cos q.
The volume of this cylinder is
6.12 Heat and Thermodynamics

r sin q rd
q

P
q q
dq

O dA

Fig. 6.14 Hemisphere around dA

dV = c dA cos q (8)
The energy density dY inside the cylinder and hence, in particular, at points
near the surface is
(dQ/dt) × 1 second
d Y = ________________
dV
IdS
= ___2 (9)
cr
Since the surface area of the hemisphere is Ú dS = 1/2 (4pr2) = 2pr2, the total
radiation density Y near the surface is given by

1 2pI
Y = Ú dY = ___2 Ú dS = ___
c (10)
cr
Using Eqn. (6), e = pI,

2e
Y = ___
c (11)

Similarly, the monochromatic energy density is


2el
Yl = ___
c (12)

The radiation from a surface is confined to a hemisphere of directions in


the front of the surface. In the closure, the radiation is black and travels in all
directions. Therefore, Ú dS = 4pr2 (or the area of the sphere) and

4e(b)
4pI ____
Y = ___
c = c ,
4e(b)
l
____
Yl = c (enclosure) (13)

in the enclosure.
Radiation 6.13

6.10 PRESSURE DUE TO RADIATION


Consider radiation falling normally on the surface of a body. With the mean
energy density w is associated the momentum density w/c. In one second all the
radiation contained in a cylinder of length c cm and cross-sectional area unity
would reach unit surface area.
Thus on each unit area of the surface in one second cw ergs of energy would
fall. This will brings w units of momentum per second, the momentum vector
being directed normally toward the surface. If the waves are absorbed by the
surface, it receives this momentum and consequently experiences a pressure
equal to w.
Suppose the radiation is incident at an angle q. The energy contained in the
cylinder of length c and cross-sectional area unity would now fall on a surface
of (larger) area 1/cos q. The force per unit area of the surface is now w cos q and
acts in the direction of rays. Its component normal to the surface is w cos2 q. If the
radiation is absorbed, the resulting pressure is
p = w cos2 q (1)
The same expression would hold for the pressure due to the emission of a beam
at an angle q. Clearly, for specular reflection at an angle of incidence q, the total
pressure on the surface would be 2w cos2 q.
Now consider the radiation in an enclosure, where radiation is moving toward
a surface and also away from it with equal intensities in all directions. Such a
distribution is equivalent to a large number N of beams, all of equal intensity.
With their directions of propagation distributed equally in direction (N/2 moving
toward the surface and N/2 moving away). Then the total energy density just in
front of the surface and the pressure p on it are given by
Y = Nw,
p = S w cos2 q = w S cos2 q (2)
2 2
where S cos q denotes the sum of the values of cos q for all of the beams.
To calculate S cos2 q, we again draw a hemisphere of radius r round a point
O at the centre of the surface element. From the hemisphere we cut out a ring of
area 2pr sin q r dq by means of two cones of semiangle q and q + dq drawn from
O as apex, as in Fig. 6.14. If dN is the number of beams that cross this ring, then
area of ring
dN ________________
___ =
N area of hemisphere
2pr sin q r dq
= ____________ = sin q dq (3)
2pr2
The value of cos2 qis the same for all of the dN beams. Therefore, their
contribution to S cos2 q is cos2 q dN or N cos2 q sin q dq. Thus
6.14 Heat and Thermodynamics

S cos2 q = Ú cos2 q dN
p/2
1
= N Ú cos2 q sin q dq = __ N (4)
0 3
and from Eqn. (2)
1 1
p = __ wN = __ Y (5)
3 3
There is an analogy between the behaviour of the radiation within an isothermal
enclosure and that of gas molecules in a box. Just as the gas molecules in a box
move with all velocities in all directions, radiation in an enclosure consists of all
frequencies and goes in all directions. The plot of the fraction of the total energy
radiated within each range of wavelength (Fig. 6.11) shows marked resemblance to
that of the Maxwell distribution of the molecules and the radiation both exchange
momentum with the walls and exert pressure on them.

6.11 STEFAN‐BOLTZMANN LAW


The radiation in an enclosure occupies a volume, exerts pressure, and is an
equilibrium with the walls at a given temperature. Therefore, it is reasonable to
apply thermodynamics to this system. Boltzmann in 1884 applied the laws of the
Carnot cycle to an engine in which the working substance is radiation.
Consider a frictionless nonconducting cylinder-piston arrangement with a
base opening O through which radiation may enter or leave, Fig. 6.15. Inside
the cylinder is a vacuum and the walls are perfectly reflecting. Such walls do not
absorb radiation and therefore the heat capacity of walls does not enter in the
calculations. With the help of an isothermal enclosure H1 at temperature T1 the
cylinder is filled up with radiation until there is the same density Y1 of radiation
inside the cylinder as there is in H1. The radiation exerts a pressure p = 1/3 Y1.

P3
P2
P4
P1 P4
p1 V2 T3
p1 V1 T1 p1 V2 T1 p2 V3 T2 p1 V4 T2

T1 H1 T2 H2

Fig. 6.15 Carnot engine for radiation


Radiation 6.15

The Carnot cycle for the radiation is run as follows:


(a) Piston moves from P1 to P2 with O open. Volume increases from V1 to V2
while the density Y1 is kept constant by supplying radiation Q1 through O
from source H1. This expansion process is isothermal.
(b) Adiabatic expansion from P2 to P3 with O covered. Volume increases and
work is done by the radiation on the piston. Both these factors cause Y1 to
decrease to Y2. The radiation pressure also falls.
(c) Isothermal compression from P3 to P4 with O open. The density of radiation
rises and radiation Q2 passes through O into sink H2. The radiation density
Y2 remains constant.
(d) Adiabatic compression P4 to P1 with O covered. Original condition is
restored.
For an infinitesimal cycle T1 = T, T2 = T – dT, Y1 – Y2 = dY and p1 – p2 = dp
= 1/3 dY, Fig. 6.16. For the isothermal expansion AB, we have
1
work done: W = p(V2 – V1) = __ Y(V2 – V1) (1)
3
change in internal energy:
U2 – U1 = Y (V2 – V1)
first law: Q1 = (U2 – U1) + W
4
= __ Y(V2 – V1) (2)
3
p

A B
p
dp
p – dp
D C

V V2 V

Fig. 6.16 Carnot cycle of the radiation engine

The net work done in the cycle is the shaded area in Fig. 6.16,
dW = dp(V2 – V1)
1
= __ dY(V2 – V1)
3
6.16 Heat and Thermodynamics

Therefore, by the usual rule of a Carnot cycle


1
__ dY(V2 – V1)
dW ____________
___ 3
=
Q1 4
__ Y(V2 – V1)
3
dT
= ___
T

dY
___ dT
= 4 ___ (3)
Y T
Integration gives ln Y = 4 ln T + constant, or
Y = aT 4 (Stefan-Boltzmann law) (4)

(We assume that Y = 0 at T = 0). Thus, the energy density is proportional to the
fourth power of the absolute temperature T.
From Eqn. (13) Section 6.9, we have e(b) = Yc/4. Therefore,
ac
e(b) = ___ T 4 = s T 4 (Stefan-Boltzmann law) (5)
4
where s is the Stefan-Boltzmann law (or, Stefan) constant. The value of s is
5.735 × 10–5 erg cm–2 sec–1 K – 4.
We can derive Eqn. (4), without bringing in the Carnot cycle, by simply
considering the isothermal compression of radiation and applying the general
laws of thermodynamics directly.
Consider radiation in a cylinder-piston arrangement. Introduce into the cylinder
a minute speck of lampblack which is kept at constant temperature T by thermal
contact with an external reservoir. For very slow compression, the radiation can
be regarded as in thermal equilibrium with the lampblack at all times. We thus
have blackbody radiation at temperature T.
The radiation has total energy density Y and exerts a pressure p = 1/3 Y. The
total internal energy of the system of volume V is
U = YV (6)
We know that the radiation is a function of temperature only,

( )
∂Y
___
∂V T
=0 (7)

From the first law dU = TdS – pdV, we have

( ) ( )
∂U
___
∂V T
∂S
= T ___
∂V T
–p
Radiation 6.17

or using U = YV and a Maxwell relation

∂Y
Y + V ___
∂V ( ) ( )
T
∂p
= T ___
∂T V
–p (8)

Since (∂Y/∂V)T = 0 and p = 1/3 Y, we get

1 ∂Y
3 ∂T ( )
Y = T __ ___
1
– __ Y
V 3
(9)

or noting that Y = Y(T),


dY
4Y = T ___ (10)
dT
Integration gives Y = const. × T 4.

6.12 VERIFICATION OF STEFAN’S LAW


Lummer and Pringsheim, and also many other workers, studied the emission from
a blackbody and verified Stefan’s law.
In the temperature range 200°C to 600°C a hollow copper sphere serves as a
black radiation source. The temperature of the bath can be varied and measured
with a thermocouple, Fig. 6.17. The hollow vessel A containing boiling water
at 100°C serves as a standard source (Fery blackbody) for the calibration of the
bolometer B.

100°C

C B A
S S¢

Fig. 6.17 Lummer and Pringsheim’s arrangement for the verification of Stefan’s law

Radiation from A is allowed to fall on B by raising the shutter S¢. The


galvanometer deflections are noted for various distances from A. If the inverse
square law is obeyed, one can assume that the bolometer deflection is proportional
to the incident radiation.
The shutter S¢ is now closed and the radiation from the source C at temperature
T is allowed to fall on B by raising the water cooled shutters S at temperature T0
(T0 < T). The steady deflection of the galvanometer is noted. The readings are
6.18 Heat and Thermodynamics

taken with the bath at different constant temperatures. These data are reduced to
a common arbitrary unit depending upon the radiation from the blackbody A at
100°C and kept at a fixed distance. By Stefan’s law

f = a (T4 – T04)

where f, the galvanometer deflection, is proportional to the excess of energy


received per second by B. The coefficient a is found to be constant for various
values of T, thereby verifying Stefan’s law. The constant a depends upon the
calibration of the bolometer and so is not Stefan’s constant s. However, the two
can be related in terms of the area of the aperture, the area of the receiver and the
distance between them.

6.13 DETERMINATION OF STEFAN’S CONSTANT


IN THE LABORATORY
The apparatus consists of a metal hemisphere M, with blackened inner surface,
enclosed in a metal steam chamber C fitted with thermometers T T, Fig. 6.18.
This system rests on a wooden platform W with a small hole in the centre. A
small silver disc D, blackened at the top, can be put in this hole to measure the
blackbody radiation from the enclosure M. A thermocouple is soldered to the
bottom of the disc D and the other junction is placed in a test tube containing

T T

C M

D
W
G

X Y

Fig. 6.18 Apparatus for measuring Stefan’s constant


Radiation 6.19

some oil and standing in a waterbath X. A


galvanometer G is connected by copper leads
to the thermocouple. The junctions of copper
q1
leads with the thermocouple wires are kept in
a vessel Y packed with cottonwool to avoid q
any disturbing effects due to the temperature
differences in the leads.
To standardize the thermocouple, M b
and D (placed in the hole) are kept at room
temperature while water in X is heated. A f1
calibration curve is plotted between the
f
galvanometer deflection f and the temperature
difference q of the two junctions, Fig. 6.19. Fig. 6.19 Standardization curve for
the thermocouple
Next the disc D is removed from the
hole, water in X is brought back to the room
temperature and steam is passed in the chamber C. When T T show steady
temperatures, the disc D is introduced in the hole. The temperature of the disc
D rises continuously at a rate depending upon the net heat gained by the disc
per second. Let H1 = s Ts4 be the amount of radiation absorbed by the disc per
unit area per second, where Ts is the steam temperature of the blackbody M. Let
H2 = s T 4 be the amount of radiation emitted by the disc per unit area per second
at the temperature T. The gain in energy per second is
dT
A(H1 – H2) = As (Ts4 – T 4) = mc ___ (1)
dt
where A is the area, m the mass, c the
specific heat, and dT/dt the rate of rise
of temperature of the disc. If T and dT/dt f
can be measured we can calculate s.
As the temperature of the disc D rises P
f1
the galvanometer deflection f increases.
A plot is obtained between f and time t,
Fig. 6.20. Consider a point P on this curve
corresponding to detection f1 at time t1.
The temperature difference q1 between μ
the disc and the room, corresponding t1 t
to the deflection f1, can be read from
Fig. 6.20 Galvanometer deflection f vs.
Fig. 6.19. Then q1 = T1 – T0 is the disc time t plot
temperature and T0 the room temperature.
Thus,
T1 = T0 + q1
6.20 Heat and Thermodynamics

To find dT/dt at t1, we draw a tangent on the t – f curve at the point P making
an angle a with the time-axis, Fig. 6.20. Then

( ___dTdt ) t = t1 [
d(T0 + q)
= ________
dt ] ( )
t = t1
dq
= ___
dt t = t1

(
dq df
= ___ ◊ ___
df dt ) ( )( )
t = t1
dq
= ___
df
df
___
dt t = t1

= tan b tan a
where in the last step we have used the fact that q-f curve is generally a straight
line making an angle b with the f-axis, Fig. 6.19. Substituting these values of T1
and (dT/dt)t1 in Eqn. (1), we can calculate s.

6.14 NEWTON’S LAW OF COOLING


If a body at temperature T is placed in a surrounding at temperature T0 < T, then
the loss of radiant heat by the body per second is given by Stefan’s law to be
H = s (T4 – T04)

= s (T – T0) (T3 + T2 T0 + T T02 + T03)


When the temperature (T – T0) is small, that is, T is nearly equal to T0, the
above result can be written as

H = s (T – T0) × 4T03
= constant × (T – T0)

6.15 ADIABATIC EXPANSION OF BLACKBODY RADIATION


For an adiabatic expansion
dQ = 0 = dU + pdV
1
Using U = VY and p = __ Y, we have
3
1
d(VY) + __ Y dV = 0
3
4 dV
dY/Y + __ ___ = 0
3 V
YV4/3 = constant (adiabatic) (1)
By Stefan’s law Y = aT4. Therefore,
T 3V = constant (adiabatic) (2)
Radiation 6.21

6.16 EFFECT OF ADIABATIC EXPANSION ON


BLACKBODY RADIATION
An enclosure at temperature T containing a small
blackbody inside it is filled up with blackbody
radiation of total density Y which depends on T
only (Stefan’s law). The blackbody is removed and
the enclosure is allowed to undergo slow adiabatic
expansion (dQ = 0, dS = 0) from volume V to V¢,
Fig. 6.21. The temperature falls to T¢ and the total
energy density reduces to Y¢. We now show that
the radiation remains black (or complete) radiation Fig. 6.21 Adiabatic expansion
of blackbody radiation
characteristics of the lower temperature T¢ if
expanded (or compressed) adiabatically in the absence of any blackbody.
Let the initial energy, entropy and volume of the radiation be Ui = Vi Yi, Si and
Vi.
(a) Expand reversibly and adiabatically from Vi to Vf doing work W(a). We
have dS(a) = 0 and dU(a) = W(a).
(b) Introduce a black speck. For the sake of argument assume that the
radiation after expansion is no longer black (or complete), that is, there
is more radiation per unit volume of wavelengths near some value l1 and
deficiency near some other wavelength l2. Consequently, an irreversible
change to blackness would take place accompanied by an increase in
entropy, dS(b) > 0. Since no work is performed and no heat enters the
system, dU(b) = 0.
(c) Keeping the speck in the enclosure, compress adiabatically and reversibly
back to Vi doing work W(c). For this process dS(c) = 0 and dU(c) = W(c). Let
the internal energy and entropy now be Uf and Sf.
Note that the pressure depends only on the total energy density Y. Therefore,
pressure is the same function of volume during the compression as it was during
the expansion, that is, W(a) = – W(c). Since no heat enters the system, Uf = Ui and
Yf = Yi. The radiation is black both initially and finally. The energy density of
black radiation depends only on the temperature (Stefan’s law). It follows that
Ti = Tf. Thus, from the facts that in the initial and the final states
(i) the volume is the same,
(ii) the radiation is black, and
(iii) the energy density is the same
we can conclude that the initial and final states are identical. Then Sf = Si and
dS(b) = 0. Therefore, there could have been no irreversible change in (b) and the
radiation must have been black after the expansion. Hence, the black radiation after
slow adiabatic expansion remains black, characteristic of the lower temperature.
6.22 Heat and Thermodynamics

6.17 WIEN’S LAW


The derivation of the law depends upon the existence of the Doppler effect which
concerns the change in wavelength when radiation is reflected from a moving
mirror or wall.
Consider radiation moving in the direction MN and being reflected in the
mirror at the position S1, which moves a distance z (position S2) with a velocity v
perpendicular to its plane Fig. 6.22. We consider the reflection of the beginning A
and the end B of the incident wave of wavelength l which is equal to the distance
AB. For the incident wave, before reflection, the frequency is v = c/l, where c is
the velocity of radiation (or light).

S1 S2
M

B
N
Q A
P
q N¢
q
q

P¢ Q¢

Z
Fig. 6.22 Doppler effect for radiation

The time taken for A to reach P is NP/c, and for B to reach P¢ is

1
__
c (l + NN¢ + N¢Q¢ + Q¢P¢) =
1
__
c l + [z
_____ +
z
_____
cos q cos q
+ NP –
cos q(
2z sin2 q
________
)]
1
= __
c (NP + l + 2z cos q)

Therefore, the time interval between the arrival of successive wave beginnings
at P and P¢ respectively is

1
__
c (l + 2z cos q)

The corresponding frequency of the reflected radiation, v¢, is equal to the


reciprocal of this time interval. Thus
Radiation 6.23

c
v¢ = __________,
l + 2z cos q
c
l¢ = __ = l + 2z cos q

Equating the time taken for the mirror to move a distance z and for the radiation
to move from B to N¢, we get
l + (z/cos q)
__z = ___________
v c
Therefore,
2lv cos q
l¢ = l + ___________ (1)
c – (v/cos q)
Since v << c, we can write for the change in wavelength
2vl cos q
dl = l¢ – l = ________
c (Doppler effect) (2)

We now consider the reflection


of radiation in a spherical
enclosure of radius r, Fig. 6.23. q
q
q
the enclosure is taken to be
spherical for convenience. The
distance between two consecutive
reflections is always 2r cos q. If the q
surface of the enclosure undergoes
slow adiabatic expansion at
uniform velocity v, the change in Fig. 6.23 Construction for the charge in wavelength
wavelength at each reflection is on reflection at the wall of an expanding enclosure
given by Eqn. (2). The number of
reflections per second being c/2r
cos q, the change of wavelength per second is

( c
dl = _______
2r cos q ) ( ________
2vl cos q
c ) = ___vlr
dl __v __
___ dr
=r= r (3)
l
where dr is the distance traversed by the surface in one second. Integration gives
l
__
r = constant (4)

For adiabatic expansion Eqn. (2) (Section 6.15) gives


6.24 Heat and Thermodynamics

(
4
TV1/3 = T __ pr3
3 ) 1/3
= constant

Tr = constant (5)
From Eqns. (4) and (5), we have
lT = constant (adiabatic) (6)
Thus, the wavelength of any given spectral component of the blackbody radiation
changes during a slow adiabatic expansion in such a way that l μ 1/T.
We now consider the effect of adiabatic expansion on the spectral components
between l and l + dl. The spectral density, the wavelength and the width of this
band will be changed by the expansion. Applying the first law (dQ = dU + pdV = 0)
again, but this time to these spectral components only, we have
d(Ul dl) + pdV = 0
1
d(Yl V dl) + __ Yl dl dV = 0
3
1
d(Yl dl)V + Yl dl dV + __ Yl dl dV = 0
3
d(Yl dl)
________ 4 dV dr
= – __ ___ = – 4 __
r
Yl dl 3 V
Yl dl r4 = constant (7)
From Eqn. (4), l is proportional to r so that Eqn. (7) becomes Yl r4 dr =
constant, or taking sum over r, n dr = r,
Yl r5 = constant (8)
Using Eqn. (5), we finally get
Yl T–5 = constant (9)
Thus, for the group of spectral components considered, we have shown
lT = constant
Yl T–5 = constant (10)
It follows that the whole distribution must obey an equation of the form
Yl = T5 f (l T) (11)
where f (l T) is an undetermined function of the product lT. We can write
Eqn. (11) in a slightly different form,
Yl = l–5 (lT)5 f (l T)
or
Yl = l–5 F (l T) (12)
Radiation 6.25

where F (l T) is another undetermined function of the product lT. This is called


Wien’s distribution law. It gives the functional form of the spectral energy density
of blackbody radiation. To go further than this is not possible with classical
concepts of thermodynamics. To find the form of F (l T), we have to introduce
new assumptions.
Since Yl is proportional to the monochromatic emissive power of a blackbody
at the same temperature, we have from Eqn. (9)
e(b)
l
T–5 = constant (13)
From Eqn. (10) and (13) we have the interesting result that from the observed
curves of the form given in Fig. 6.11, one can obtain a single curve by plotting
Yl/T5 or e(b)
l
/T 5 against the product lT, Fig. 6.24. If the theory is true, this curve
would be the same at all temperatures. The theoretical prediction is in agreement
with the observations. It follows that if lm is the wavelength at which Yl and e(b) l
have their maximum value,
lm T = constant (for all temperatures) (14)

x
x
x x

(b) x
x
el x
5 x
T x x
x

lT

Fig. 6.24 Composite curve for the experimental verification of Wien’s law.
∫ T = 1646°K, × ∫ T = 1449°K, ∫ T = 1259°K.

This is Wien’s displacement law. The value of constant is 0.29 cm°K. For the
sum if we take T = 6000°K, we get lm = 5000Å in the visible region. For bodies at
room temperature the radiation is most intense at lm ~ 105 Å in the far infrared.

6.18 WIEN’S FORMULA


To obtain a form for F(lT) Wien in 1896 introduced some special arbitrary
assumptions regarding the nature of radiation. He assumed that the radiation inside
the enclosure is produced by some kind of resonators of molecular dimensions
and that the frequency of the waves emitted by the resonators is proportional to
their kinetic energies. That is
6.26 Heat and Thermodynamics

1 2
__ ac
mv = av = ___ (1)
2 l
where a is a constant and v is the frequency of the radiation emitted. He further
assumed that the energy (or speed) distribution of the resonators obeys the law of
Maxwell and Boltzmann. That is, the number of resonators with energies between
av and a(v + dv) is given by

( )
mv2
Nv = f(v) exp – ____
2kT
av
( )
= f(v) exp – ___
kT
(2)

If the energy density is taken to be proportional to Nv, we get


av
( )
Yv dv = Af(v) exp – ___ dv
kT
(3)

where A is a constant. Since v = c/l, we get


dv = – (c/l2) dl (4)
ac
(
Yl dl = Af(l) exp – ____ dl
klT ) (5)

Comparison with Wien’s law (Eqn. 12 of Section 6.17) gives


1
f¢(l) = __5 (6)
l
b
( )
F(lT) = constant × exp – ___ ,
lT
b = ac/k (7)
const.
Yl dl = _____
l 5 ( )
b
exp – ___ (Wien formula)
lT
(8)

The Wien formula (8) with two underdetermined constants fits the experimental
data in the short wavelength region but shows disagreement in the long wavelength
region.

6.19 RAYLEIGH‐JEANS FORMULA


On the basis of the law of equipartition of energy we have shown that for a vibrator
(or oscillator) the average energy for each vibrational degree of freedom is kT

(1
__
2
1
)
kT for the kinetic energy plus __ kT for the potential energy . Lord Rayleigh
2
(1900) and Jeans (1905) suggested that the energy equipartition law can be applied
to the radiation problem.
Radiation 6.27

The radiation within an enclosure can be considered to consist of stationary


waves of various frequencies. In the one-dimensional case of vibrating string
(Fig. 6.25), the wavelength l of a standing wave is such that
l
l = n __, n = 1, 2, 3, …, • (1)
2

l
2

Fig. 6.25 Vibrating string

Thus, the ‘allowed’ wavelengths are


2l
l = __
n, n = 1, 2, 3, …, • (2)

In the three-dimensional case we can calculate the allowed modes of vibrations


by considering stationary waves in a hollow cube of side l, with the walls of the
cube as nodal planes. For one direction then
l = nx l/2
If the vibrations propagate making an a with the direction, we have to replace
l by the projection of cube-edge on the direction of propagation. That is
l cos a = nx l/2 (3)
In general, if the angles made with the three edges forming the three axes in
space are a, b, g, then
l l l
l cos a = nx __, l cos b = ny __, l cos g = nz __ (4)
2 2 2
where nx, ny, and nz are
_› positive integers.
Consider a vector l forming the angles a, b, g with the three axes, Fig. 6.26.
Its three components have the lengths
lx = l cos a, ly = l cos b, lz = l cos g
Since l2 = lx2 + ly2 + lz2, the direction cosines must obey the relation
cos2 a + cos2 b + cos2 g = 1 (5)
From Eqns. (4) and (5), we get

( )
2l
nx2 + ny2 + nz2 = __
l
2
(6)
6.28 Heat and Thermodynamics

rh

a b
y

Fig. 6.26 Direction cosines

Thus, allowed modes of vibration are those for which (6) holds for every set of
three positive integers nx, ny, nz.
If we put 2l/l = n, Eqn. (6) becomes nx2 + ny2 + nz2 = n2. Then n is the radius
of a sphere in (nx, ny, nz) space. The number of allowed modes can be evaluated
by making a three-dimensional plot of nx, ny, and nz (Fig. 6.27) and counting the
number of points, for which these numbers are integral, which lies within the
positive octant of a sphere of radius n. For sufficiently large values of n, each
point will correspond to one unit cube (each point is shared by adjacent cubes)
in this octant. Therefore, the volume of the octant gives the required number of
modes of vibration g as

(
1 4
8 3 )14 2l
g = __ __ pn3 = __ __ p __
83 l ( ) 3

4p l3
= _____ (7)
3l3
The number of modes, or of degrees of freedom, between l and l + dl is
obtained by differentiating Eqn. (7). Thus

4pV
dg = ____ dl (8)
l4
where V = l3 and both dg and dl are taken as positive quantities for simplicity.
This number should be multiplied by 2 because transverse electromagnetic waves
have two possible polarizations for each mode. Thus, the number of degrees of
freedom of the waves is

8pV 8pV 2
dg = ____
4
dl = ____ v dv (9)
l c3
The law of equipartition of energy allows us to associate an energy kT with
each degree of freedom. Therefore, the total energy inside the box within the
wavelength range l and l + dl is
Radiation 6.29

nx
6
n

5
4
n
3

2
1
1 2 3 4 5 6
0 ny
1 n
2
3
4
5
6
n
nx

Fig. 6.27 Points in the (nx, ny, nz) space

8pV
VYl dl = dg kT = ____
l4
dl kT ( )
or
8pkT
Yl dl = _____ dl
l4

8pk
= ____ lT dl (Rayleigh-Jeans law) (10)
l5

8pv2
Yv dv = ____ kT dv
c3
This is the Rayleigh-Jeans law, which results from the application of
the classical equipartition law. In this case F(lT) = lT. This law agrees with
observations at long wavelengths but fails miserably at short wavelengths or high
frequencies. For example, as l Æ 0, Yl Æ •, when we know (Fig. 6.11) that
actually as l Æ 0, Yl Æ 0.

The total energy is given by Ú Yv dv. With Yv given by the Rayleigh-Jeans


0
law, this integral gives infinite value (diverges) because Yv increases with v.
It means that the energy density of blackbody radiation, when in equilibrium
6.30 Heat and Thermodynamics

with the walls of a enclosure, is infinitely greater than the energy density in the
walls themselves. This is not in agreement with the observations and is called
the ultraviolet catastrophe. Let us see what will happen if, for example, we fill
an enclosure with red light. The radiation will not be in equilibrium with the
enclosure. As exchange of energy with the walls brings about the equilibrium,
the energy distribution will approach the Rayleigh-Jeans distribution; that is, the
energy will tend to get concentrated in the region of very short wavelengths. This
implies that the red light will turn into violet, ultraviolet, X-rays and g-rays, etc.
This is contrary to all observations.

6.20 PLANCK’S FORMULA


The Wien formula was found to fit the experimental curve in the small wavelength
region and the Rayleigh-Jeans formula in the very long wavelength region.
Otherwise, both formulas failed. The failure of Rayleigh-Jeans formula showed
a serious limitation of classical physics because it involved no underdetermined
constants and the derivation was flawless.
The essential failure of the Rayleigh-Jeans formula was that it led to the
0 •

ultraviolet catastrophe. The total energy density Y = Ú Yl dl = Ú Yv dv diverges


• 0

(becomes infinite) as l Æ 0 or v Æ •, contrary to observations. The only way to


avoid this is to discourage the radiation oscillators from exchanging energy with
the enclosure walls as l becomes small or v becomes large. This is easily achieved
by assuming with Max Planck (1900) that the radiation oscillators exchange
energy with their surroundings not continuously (as in classical physics) but in
discrete units
en = n h v (1)
–27
where h is Planck’s constant (6.6 × 10 erg sec.) and n is an integer. The available
states are only those for which energy is 0, hv, 2hv, …, etc. Now when v is large,
the available states are widely separated in energy and the radiation oscillator
can reach it only by the absorption of very high energy quanta in one step, a very

rare occurrence. Thus, no contribution to the integral Y = Ú Yv dv is obtained for
0
v Æ • and the ultraviolet divergence or catastrophe is avoided.
Consider a collection of N oscillators having a characteristic vibration frequency
v. By Planck’s assumption these can absorb energy only in increments of hv. If N0
is the number of oscillators in the lowest energy state, the number Nn, in a state
whose energy is en, is given by the Boltzmann distribution law
Nn = N0 exp (– en/kT), en = n h v (2)
Radiation 6.31

The total number N of the oscillators is the sum of all Nns from n = 0 to
n = •,
N = N0 + N1 + N2 + N3 + …
= N0 + N0 exp (– hv/kT) + N0 exp (– 2hv/kT) + …

= N0 S exp (– nhv/kT) (3)
n=0

The total energy is the sum of Nn en,



U = N0 S nhv exp (– nhv/kT) (4)
n=0
_
Therefore, on the Planck’s hypothesis, the average energy e of an oscillator is

_ U hv S n exp (– nhv/kT)
e = __ = __________________
N S exp (– nhv/kT)
Snyn
= hv ____
Syn
, ( x = ___kThv , y = e )
–x

y(1 + 2y + 3y2 + …)
_________________
= hv
1 + y + y2 + …
y/(1 – y)2 y
= hv ________ = hv _____
1/(1 – y) 1–y
hv
= ______
1/y – 1
hv
= _____________ (5)
exp (hv/kT) – 1
We can now combine Planck’s expression for the average energy of an oscillator
with the Rayleigh-Jeans analysis of the electromagnetic radiation where the
various modes of vibration were treated as oscillators. There were 8pdl/l4 modes
of vibration or degrees of freedom per unit volume in the wavelength region l
to l + dl. The Rayleigh-Jeans formula was obtained by multiplying this by kT
according to the classical energy equipartition law, which applies only when the
modes of oscillation are continuously distributed as to energy. _
In the quantum
theory of Planck we multiply this number not by kT but by e as given by (5).
Thus
dl hv
Yl dl = 8p ___4 ◊ _____________
l exp (hv/kT) – 1
8p ch ______________
dl
= _____
5
◊ (Planck’s formula) (6)
l exp (ch/lkT) – 1
6.32 Heat and Thermodynamics

Ray
leig
h-J
ea
Planck’s

ns
law

’ la
Yl

w
Wien’s law

0 1 2 3 4 5 6
4
l (cm × 10 )

Fig. 6.28 Comparison of the three radiation laws with experiment at 1600°K. The ordinate
represents Yl or e(b)
l
on an arbitrary scale

This is the famous Planck radiation law. It is in complete agreement with the
experiments of Rubens and Kurlbaum in the entire wavelength range, Fig. 6.28.
Clearly,
c F(lT)
e(b) = __ Yl = _____ ,
l 4 l5
2pch
F(lT) = ______________ (7)
exp (ch/lkT) – 1
To fit the experimental data, we need
h = 6.61 × 10–27 erg sec
If the values of constants c, h and k are inserted, Eqn. (6) is
4.98 × 10–15 _________________
1 ergs
Yl = __________ ◊ ____ (8)
l 5 exp (– 1.435/lT) – 1 cm3
where l is in centimeters and T in degrees Celsius.
The Planck formula reduces to Wien’s formula for shorter wavelengths,
exp (ch/lkT) >> 1 for lT << 1
8pch
_____
Y(Planck)
l
exp (– ch/lkT)
l5lT << 1

and to the Rayleigh-Jeans formula for longer wavelengths,


ch c2h2
exp (ch/lkT) = 1 + ____ + _______ +…
lkT 2l2k2T 2
Radiation 6.33

ch
= 1 + ____ for lT >> 1
lkT
lkT _____
8pch ____ 8pkT
Y(Planck)
l
Æ _____ 5 ◊ = 4
lT >> 1 l ch l

6.21 DETERMINATION OF THE SOLAR CONSTANT


The fraction dI/I of the incident radiation I absorbed is experimentally found to be
proportional to the thickness of the medium dz that is traversed. Thus
– (dI/I) = m dz (1)
where the constant m is called the absorption coefficient. Integration gives the
Lambert law
I = I0 exp (– mz) (2)
where I0 is the value of I at z = 0.
The solar radiation traverses the
atmosphere before it reaches the earth’s
surface. The absorption by the atmosphere
must therefore be taken into account.
When the zenith angle is b (angle
made by the sun from the vertical), the
b

H
ec

length of the atmosphere traversed by the


Hs

radiation before it reaches earth is H sec


b, where H is the height of ‘equivalent’ 90 – b
constant density atmosphere, Fig. 6.29. Horizon
The density of real atmosphere changes Fig. 6.29 Equivalent constant density
with height. Therefore, it is replaced by atmosphere and zenith angle
an equivalent atmosphere of height H and
constant density. We thus have
I = I0 exp (– m H sec b)
= I0 [exp (– m H)]sec b
= I0 asec b, a = exp (– m H) (3)
where the constant a is called the transmission coefficient.
In the absence of the atmosphere the mean value of solar constant would be
S0 = 60I0. The corresponding quantity in the presence of atmosphere is given by
S = S0 asec b (4)
log S = (log a) sec b + log S0
6.34 Heat and Thermodynamics

The measurements with the help of a


pyrheliometer of known exposure area yield values log S

of S for different values of b. We can plot a straight


line graph between log S and sec b, Fig. 6.30. The
intercept of this line on the log S-axis gives the
value log S0. Thus, S0 is determined. Its value is
log So
about 1.93 cal cm–2 minute–1 or about 1.34 × 106
ergs cm–2 sec–1.
O sec b
6.22 TEMPERATURE OF THE SUN
Fig. 6.30 Log S vs. sec b graph
Knowledge of S0 enables us to estimate the
temperature of the sun. The total energy radiated per second from the surface of
the sun is
S0
Q = 4pR02 ___ (1)
60
If r is the radius of the sun, then the total radiating surface area is 4pr2 and
energy emitted per second per unit area is
Q R02 S0
____ = ___ ◊ ___ (2)
4pr2 r2 60
By Stefan’s law we must have
Q
____ = s T4 (3)
4pr2
where T is the surface temperature of the sun. Thus
R02 S0 1/4
(
T = ____2 ◊ ___
s r 60 ) (4)

Here r/R0 is the mean angular radius of the sun at the earth (angle subtended
by the solar radius at the earth). Using
S0 = 1.937 cal cm–2 minute–1
s = 5.735 × 10–5 erg sec–1 cm–2 °K–1
r/R0 = 959≤ = 4.65 × 10–3 radians
we get
1.937 × 4.18 × 10–7
( )
1/4
T = ___________________________
60 × (4.65)2 × 10–6 × 5.735 × 10–5

= 5742°K
where we have multiplied S0 by J = 4.18 × 107 ergs/cal to express it in ergs. This
is the brightness temperature of the sun’s surface.
Radiation 6.35

The temperature of the sun can also be estimated by using Wien’s law
lmT = constant = 0.293
Abbot’s measurements show that lm = 4753Å for the solar radiation. Hence
0.293
T = __________ = 6060°K
7553 × 10–8
This temperature, called the colour temperature, is nearly 300°K higher
than the brightness temperature. This shows that the sun does not radiate like a
blackbody.
Example The rate of cooling or warming is given by Newton’s law of cooling
as
d
__ (DT) = – KDT,
dt
where K is a constant. On what factors does K depend and what are its dimensions?
If at some instant, the temperature difference is DT0, show that at a time t later
DT = DT0 e–kT.
Solution The constant K depends on the nature of the exposed surface and its
area. The dimension of K is (Time)–1 as time enters the denominator of l.h.s. of
the equation (and temperature is in the numerator of both sides).
Now,
d
__ (DT) = – KDT
dt
d(DT)
_____
or = – Kdt
DT
Integrating
ln DT = – Kt + C
where C is the constant of integration.
Let DT = DT0 at t = 0.
\ ln DT0 = C
and
ln DT = – Kt + ln DT0
or ln DT – ln DT0 = – Kt

or
( )
DT
ln ____ = – Kt
DT0
or DT = DT0 e–Kt
at a later time t.
6.36 Heat and Thermodynamics

QUESTIONS
1. What is meant by a blackbody? On what factors does the blackbody
radiation depend and how?
2. Explain the meaning of emissive power and absorptive power. State the
Kirchhoff’s law of blackbody radiation and explain its significance.
3. What is meant by the pressure of radiation? Show that the pressure of
radiation of diffused radiations enclosed inside an enclosure is equal to
1/3rd of energy density.
4. Explain Stefan-Boltzmann’s law and prove it thermodynamically. Show
that the Newton’s law of cooling is an approximation of it for low
temperature difference.
5. State the Stefan’s law of radiation. Describe the experimental method of
determination of Stefan’s constant.
6. What is solar constant? How can the surface temperature of sun be
determined with its help?
7. State Wien’s distribution law and derive it.
8. For an adiabatic expansion of blackbody radiation in an enclosure, prove
that TV1/3 = constant.
9. Calculate the number of modes of vibrations per unit volume of an enclosure
in the frequency range v and v + dv. On this basis, derive Rayleigh-Jean’s
law and discuss its drawbacks.
10. What is Planck’s quantum hypothesis of radiation? Obtain Planck’s
formula for distribution of energy in the blackbody spectrum.
11. Show that the average energy of a Planck’s oscillator of frequency v in
thermal equilibrium at an absolute temperature T is given as
__ hv
E = ________
hv/kT
e –1
Hence show that as h tends to zero, the average energy of Planck’s oscillator
becomes independent of frequency like the average energy of a classical
oscillator.
12. Deduce (i) Wien’s law, (ii) Rayleigh-Jean’s law from the Planck’s
distribution law.
13. Write short notes on
(i) Kirchoff’s law of radiation
(ii) Prevost’s theory of heat exchange
(iii) Stefan’s law
(iv) Wien’s displacement law
Radiation 6.37

(v) Wien’s law


(vi) Rayleigh-Jean’s law
(vii) Energy of Planck’s oscillator
(viii) Planck’s law of radiation

OBJECTIVE TYPE QUESTIONS


1. The transfer of heat energy by some sort of collision without gross
movement of any part of the body due to temperature difference between
the parts of the body is called
(a) conduction (b) radiation
(c) convection (d) none of these
2. Electromagnetic radiations are emitted from
(a) only from the red hot body
(b) the antenna of radio and TV
(c) only the body at a high temperature present in the surrounding
(d) all the above bodies
3. The speed of heat radiations is
(a) 300 ms–1 (b) 3 × 108 cm s–1
(c) 3 × 108 ms–1 (d) 3 × 10–10 ms–1
4. The intensity of radiation measured by a thermopile placed at a distance d
from the source is I. If the distance d of thermopile is doubled, the intensity
of radiation will be
I I
(a) __ (b) __
2 4
(c) 2I (d) 4I
5. A perfectly blackbody is
(a) good absorber but bad reflector of heat radiation
(b) good absorber and good reflector of heat radiation
(c) only the good reflector of heat radiation
(d) none of these
6. For a perfectly blackbody, the absorption power is
(a) 1 (b) more than 1
(c) less than 1 (d) zero
7. Stefan Boltzmann’s law is
(a) E = sT 5 (b) lmT = constant
4 4
(c) E = s(T – T0 ) (d) E = sT 4
6.38 Heat and Thermodynamics

8. The value of solar constant is


(a) 18.6 × 108 Jm–2 minute–2 (b) 2.04 × 104 Jm–2 minute–1
(c) 8.4 × 104 Jm–2 minute–1 (d) none of these
9. Wien’s displacement law is
1
(a) lT = constant (b) __ l/T = constant
3
l lT
(c) ___ = constant (d) ___ = constant
TV V
10. In the spectrum of blackbody radiation, the energy distribution can be
completely explained by
(a) Wien’s law (b) Planck’s law
(c) Stefan’s law (d) Rayleigh-Jeans law
11. The number of allowed vibration modes per unit volume in the wavelength
range l and l + dl is
8p 8pl2 dl
(a) ___4 dl (b) _______
l c3
4pl2 dl 4p
(c) _______
3
(d) ___4 dl
c l
12. According to Planck’s hypothesis, the adsorption or emission of energy
is
(a) zero
(b) only hv
(c) continuous
(d) discrete in integral multiple of hv
13. The average energy of Planck’s oscillator at an absolute temperature T and
frequency v is
(a) kT (b) hv
hv
(c) nhv (d) ________
hv/kT
e –1
Answers
1. (a); 2. (d); 3. (c); 4. (b); 5. (a); 6. (a); 7. (c);
8. (c); 9. (a); 10. (b); 11. (a); 12. (d); 13. (d).
References

1. Mark W. Zemansky and Richard H. Dittman, Heat and Thermodynamics,


McGraw Hill, Inc. (1981).
2. Francis W. Sears and Gerhard L. Salinger, Thermodynamics, Kinetic
Theory and Statistical Thermodynamics, Narosa, New Delhi (1986).
3. Richard E. Sonntag, Claus Borgnakke and Gordon J. Van Wylen,
Fundamentals of Thermodynamics, John Wiley & Sons, Inc. (2003).
4. G. Venkataraman, A Hot Story, Universities Press (India) Ltd. (1993).
5. B.K. Agarwal, Thermal Physics, Lokbharti Publications, Allahabad (India)
(1988).
6. A.K. Saxena, An Introduction to Thermodynamics and Statistical
Mechanics, Narosa, New Delhi (2010).
7. R.C. Srivastava, Subit K. Saha and Abhay K. Jain Thermodynamics, A
Core Course, Prentice Hall of India, New Delhi (2007).
8. S.K. Roy, Thermal Physics and Statistical Mechanics, New Age
International, New Delhi (2001).
Index

A Clausius inequality 5.26


Absolute 1.6 Clausius statement 5.5
Absolute temperature scale 5.17 Clausius theorem 5.24, 5.27
Absorption coefficient 6.6, 6.33 Closed system 1.18
Adiabatic 3.43 Coefficient of linear expansion 1.30
Adiabatic demagnetization 3.30, 3.41, Coefficient of performance 5.14
3.43 Coefficient of pressure 4.35
Adiabatic expansion of gases 4.12 Coefficient of viscosity 2.41
Adiabatic process 2.13 Coefficient of volume expansion 2.37
Collision cross section 1.24
B
Compression (or expansion) ratio 5.39
Blackbody 6.5, 6.9
Conduction 6.1
Bolometer 6.3
Constant volume gas thermometer 3.35
Boyle’s Law 1.12, 1.21
Continuity of the liquid and gaseous states
Boyle temperature 3.13 4.30
British thermal unit 1.28 Convection 6.1
Brownian motion 2.1 Covolume 5.58
Btu 1.27 CP – CV for a perfect gas 5.58

C D
Calorie 1.27 Dalton’s Law of partial pressures 2.15
Carnot cycle 5.8 Degrees of freedom 2.17
Carnot engine 5.7 Differential air thermometer 6.2
Carnot steam engine 5.34
Cascade process 3.35 E
Celsius scale 1.2 Efficiency of carnot engine 5.11, 5.15
Centigrade scale 1.3 Efficiency of engine 5.7
Charle’s law 1.13, 2.12 Ehrenfest equations 5.67
Claude process 3.39 Einstein’s theory of translational brownian
Clausius-clapeyron equation 5.63 motion 2.4
Clausius-clapeyron latent heat equation Enthalpy 2.20, 5.46
5.61 Entropy 5.27
I.2 Index

Equivalence of the thermodynamic scale Joule-Thomson coefficient 3.24


with the perfect gas scale 5.20 Joule-Thomson expansion 3.43
Joule-Thomson inversion temperature
F 3.23
Fahrenheit scale 1.2 Joule-Thomson’s effect 3.17
First law of thermodynamics 4.5, 4.14
First TdS equation 5.56 K
Kamerlingh-Onnes method 3.40
G Kelvin scale 1.6
Gases 3.10 Kelvin statement 5.4
Gas thermometer 3.44 Kinetic theory of gases 2.8
Gibb’s free energy 5.47 Kirchhoff’s law 6.8

H L
Heat capacity 1.26, 4.15 Langevin’s theory of Brownian Motion
Heat capacity at constant pressure 1.26 2.2
Heat capacity at constant volume 1.26 Law of corresponding states 3.15
Heat engine 5.6 Linde process 3.39
Heat waves 6.1 Liquefaction 3.4
Helmholtz free energy 5.45, 5.46 Liquid thermometer 3.44

I M
Indicator diagram 4.3 Magnetic thermometer 3.45
Infrared radiations 6.2 Magnetocaloric effect 3.30
Intensive 4.6 Maxwell Boltzmann law of distribution of
Internal energy 4.7 velocities 2.36
Inversion curve 5.1 Maxwell Boltzmann’s speed distribution
Irreversible process 1.20 function 1.27
Isobaric 1.18 Maxwell’s relations 5.52
Isobaric processes 1.18 Monochromatic emissive power 6.6
Isochoric 4.25
N
Isochoric processes 1.20, 4.11
Negative temperatures 5.43
Isotherm 4.35
Newton’s law of cooling 6.20
Isothermal 4.11
J O
Joule’s coefficient 3.17 Open system 1.18
Joule’s expansion 3.15 Otto cycle 5.37
Joule’s expansion of an Ideal Gas 3.15
Joule’s expansion of van der Waal’s or P
Real Gas 3.16 Partial derivative 4.32
Joule’s law 4.36 Perfect gas 1.12, 4.37
Index I.3

Perfect gas equation 2.14 State function 1.20


Perfect (or ideal) gas 3.2 State variables 1.17
Perfect reflector 6.5 Stefan-Boltzmann law 6.16
Perpetual-motion machine 5.4 Stefan constant 6.16
Phase change of first order 5.27 Stern’s experiment 1.17
Physical significance of entropy 3.18
Planck radiation law 6.32 T
Planck’s constant 6.30 TdS equations 5.56
Pressure law 5.41 Temperature 1.3
Prevost’s theory of exchanges 6.6 Temperature of inversion 3.19
Principle of degradation of energy 4.9 Theorem of maximum work 5.43
Thermal conductivity coefficient 2.42
Q Thermocouple 1.11
Quasi-static 4.2 Thermocouple thermometer 3.44
Quasistatic irreversible process 5.2 Thermodynamic coordinates 1.19
Quasistatic process 1.20 Thermodynamic potentials 5.44
Thermodynamic system 1.19
R Thermopile 6.3
Radiation 6.1 Third law of thermodynamics 5.68
Rankine cycle 5.35 Throttling process 3.17
Rayleigh-Jeans law 6.29 Torr 1.7
Real gas 3.1 Total differential 4.33
Reciprocal theorem 4.34 Total emissive power 6.6
Reciprocity theorem 3.43 Total emissivity 6.9
Regenerative cooling 3.37 Total enthalpy 4.15
Regenerative cooling method 4.17
Relationship Between Cp and Cv 3.44 U
Resistance thermometer 2.28, 5.2 Ultraviolet catastrophe 6.30
Ritchie’s experiment 6.5 Universal gas constant 1.16

S V
Second law of thermodynamics 5.4 van der Waal forces 3.15
Second order phase transition 5.67 van der Waals constants 3.10
Second TdS equation 5.56 van der Waals equation 3.8
Specific 4.15 van der Waals equation of state 1.24
Specific enthalpy 1.19 van der Waals force 1.23
Specific heat capacity 1.3 van der Waals interaction 1.22
Specific value 1.7 Vapour 3.6, 3.10
Standard atmosphere 4.34 Vapour pressure 3.6
Standard fixed point of thermometry 2.33 Vapour pressure thermometer 3.44
I.4 Index

Variation of specific heat of gases with Wien’s displacement law 6.25


temperature 4.22 Wien’s distribution law 6.25
VAT – VUS diagram 5.54
Virial coefficients 3.7 Z
Zartman and KO’s experiment 2.34
W Zeroeth law 1.1
Wien formula 6.26 Zero 1.6

You might also like