100% found this document useful (1 vote)
252 views

Lecture Notes EME 2315 (Original)

This document is a lecture note on thermodynamics for a 5th year BSc program in Telecommunication and Information Engineering. It covers key concepts in thermodynamics including properties of substances, the first and second laws of thermodynamics, working fluids, thermodynamic cycles like Rankine cycle, and heat transfer modes like conduction, convection and radiation. The document contains chapter outlines, definitions of technical terms, equations, and diagrams to illustrate thermodynamic concepts and processes.

Uploaded by

Hosea Muchiri
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
252 views

Lecture Notes EME 2315 (Original)

This document is a lecture note on thermodynamics for a 5th year BSc program in Telecommunication and Information Engineering. It covers key concepts in thermodynamics including properties of substances, the first and second laws of thermodynamics, working fluids, thermodynamic cycles like Rankine cycle, and heat transfer modes like conduction, convection and radiation. The document contains chapter outlines, definitions of technical terms, equations, and diagrams to illustrate thermodynamic concepts and processes.

Uploaded by

Hosea Muchiri
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 77

UNIT CODE: EME 2315

UNIT TITLE: THERMODYNAMICS

5th Yr BSc Telecommunication and Information


Engineering, DKU T Lecture Notes

Anthony Gı̃tahi

January - April, 2012


TABLE OF CONTENTS

Preamble . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.0 Introduction 3
1.1 What is thermodynamics? . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Terminologies in thermodynamics . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Substance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.2 System, boundary and surroundings . . . . . . . . . . . . . . . . 4
1.2.3 Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.4 Process, state and property . . . . . . . . . . . . . . . . . . . . 5
1.2.5 Cycle and path . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.6 Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.7 Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 The Zeroth Law of Thermodynamics . . . . . . . . . . . . . . . . . . . 7
1.4 First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.1 Closed system - Non-flow energy equation . . . . . . . . . . . . 8
1.4.2 Open system - Steady flow energy equation . . . . . . . . . . . . 9

2.0 The Working Fluid 13


2.1 Liquid, vapour and gas . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Saturation state . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.2 Perfect gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 Vapour tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.1 Wet vapour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.2 Superheated vapour . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.3 Interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Thermodynamic charts/diagrams . . . . . . . . . . . . . . . . . . . . . 19
2.3.1 T-s diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.2 Mollier Chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.0 The Second Law Of Thermodynamics 23


3.1 The heat engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Reversibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 Deriving the second law of thermodynamics . . . . . . . . . . . . . . . 26
3.4 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.5 The Carnot cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

i
3.6 Absolute temperature scale . . . . . . . . . . . . . . . . . . . . . . . . . 33

4.0 Vapour Power Cycles 34


4.1 The Rankine cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2 Rankine cycle with superheat . . . . . . . . . . . . . . . . . . . . . . . 35
4.3 Isentropic efficiencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3.1 Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3.2 Pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.4 Steam Turbines in Cogeneration . . . . . . . . . . . . . . . . . . . . . . 39

5.0 Heat Transfer 43


5.1 Conduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.1.1 Multilayer wall . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.1.2 Hollow cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.1.3 Composite cylinder . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.1.4 The heat diffusion equation . . . . . . . . . . . . . . . . . . . . 48
5.2 Convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.2.1 Overall heat transfer coefficient . . . . . . . . . . . . . . . . . . 53
5.3 Critical thickness of insulation . . . . . . . . . . . . . . . . . . . . . . . 56
5.4 Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4.2 Blackbody and Graybody radiation . . . . . . . . . . . . . . . . 59
5.4.3 Spectral distribution of blackbody radiation . . . . . . . . . . . 60
5.5 Heat Exchangers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.5.1 Types of Heat Exchangers . . . . . . . . . . . . . . . . . . . . . 63
5.5.2 Types of Recuperative Heat Exchangers . . . . . . . . . . . . . 64
5.5.3 Effectiveness-NTU Method . . . . . . . . . . . . . . . . . . . . . 69

ii
LIST OF FIGURES

1.1 Power plant: closed system . . . . . . . . . . . . . . . . . . . . . . . . 4


1.2 Power plant: open system . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Open system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1 Liquid line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13


2.2 Vapour line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Combined line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.5 Outline of a T-s diagram for steam . . . . . . . . . . . . . . . . . . . 19
2.6 Outline of Mollier chart for steam . . . . . . . . . . . . . . . . . . . . 21
2.7 Mollier chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.1 Heat engine. The gas inside the cylinder is the system . . . . . . . . . 23
3.2 pv diagram for the heat engine . . . . . . . . . . . . . . . . . . . . . . 24
3.3 (a)Reversible and (b)Irreversible work transfer . . . . . . . . . . . . . 26
3.4 (a)Reversible and (b)Irreversible heat transfer . . . . . . . . . . . . . 27
3.5 Heat engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.6 Reversible and irreversible heat engine . . . . . . . . . . . . . . . . . 28
3.7 The Carnot engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.8 Carnot T-s cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.9 A T-s diagram for a Carnot cycle using steam . . . . . . . . . . . . . 32

4.1 T-s diagram for the Rankine cycle . . . . . . . . . . . . . . . . . . . . 35


4.2 Rankine cycle with superheat . . . . . . . . . . . . . . . . . . . . . . 36
4.3 Isentropic efficiencies . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4 An ideal steam-turbine cogeneration plant . . . . . . . . . . . . . . . 40
4.5 A practical cogeneration plant . . . . . . . . . . . . . . . . . . . . . . 41
4.6 Schematic diagram for the cogeneration plant . . . . . . . . . . . . . 42

5.1 One-dimensional conduction . . . . . . . . . . . . . . . . . . . . . . . 44


5.2 Multilayer plane wall . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3 Hollow cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.4 Composite cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.5 Heat diffusion equation-Cartesian . . . . . . . . . . . . . . . . . . . . 49
5.6 Heat diffusion equation-Cylindrical . . . . . . . . . . . . . . . . . . . 51

iii
5.7 Composite wall - overall heat transfer coefficient . . . . . . . . . . . . 54
5.8 Composite cylinder with convection . . . . . . . . . . . . . . . . . . . 55
5.9 Pipe insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.10 Reflectivity, absorptivity and transmissivity of incident radiation . . . 59
5.11 Planck distribution for various temperatures . . . . . . . . . . . . . . 61
5.12 Variations of the fluid temperatures for (a) parallel flow (b) counter
flow heat exchangers . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.13 NTU relations for heat exchangers . . . . . . . . . . . . . . . . . . . . 71

iv
PREAMBLE
Purpose
The aim of this course is to enable the students to:

1. Understand the principles of energy conservation and efficiency of conversion of


heat into work.

2. Understand the properties of working fluids commonly used in thermodynamic


processes.

3. Understand the operation of typical power cycles.

4. Learn the principles of the various modes of heat transfer and application in heat
exchangers.

Course Objectives
At the end of this course, the students should be able to:

1. Apply the first and second law of thermodynamics to typical closed and open
processes

2. Analyze vapour power cycles.

3. Apply laws governing heat conduction to simple solid geometries.

4. Design simple heat transfer devices and select heat exchangers.

Course Description
Scope of classical thermodynamics: The concept of the zeroth law of thermody-
namics. Concept of state functions. Work, heat, internal energy and enthalpy.
First law of thermodynamics: Steady-flow energy equations and application to
boilers, condensers and turbines.
Second law of thermodynamics: The concept of heat engine. Entropy.
Vapour power cycles: Rankine cycle. Thermodynamic properties of steam (steam
tables and Mollier diagram). Steam turbines; performance of steam turbines (temperature-
entropy). Application of steam turbines to co-generation. Pressure turbines.
Heat Transfer: Modes of heat transfer. One-dimensional and two-dimensional steady
state conduction. Insulation. Heat transfer by convection: Natural and forced convec-
tion. Radiation. Heat exchangers: Types and determination of effectiveness.

1
Course Textbook(s)
1. G. F. C. Rogers & Y. R. Mayhew (1992) Engineering Thermodynamics, 4th
Edition.

2. Holman, J. P. (2002) Heat Transfer, McGraw Hill, 9th Edition.

3. Michael J. Moran and Howard N. Shapiro (2003) Fundamentals of Engineering


Thermodynamics, Wiley, 5th Edition.

4. Irving Granet & Maurice Bluestein (2003) Thermodynamics and Heat Power,
Prentice Hall, 7th Edition.

References
1. T. D. Eastop & A. McConkey (1993) Applied Thermodynamics for Engineering
Technologists, 5th Edition.

2. M. D. Burghardt (1993) Engineering Thermodynamics, Harper Collins.

3. Lynn D. R. & George A. A. (1993) Classical Thermodynamics, Oxford University


Press.

4. Frank P. I. & David P. D. (1990) Introduction to Heat Transfer, John Wiley &
Sons, 3rd Edition.

2
1.0 Introduction
1.1 What is thermodynamics?
In today’s world, we need to convert the various sources of energy e.g. coal, oil, natural
gas, wind, solar etc into useful mechanical form.
Secondly, some of these forms of energy are non-renewable and are being depleted at
a fast rate. The world is then faced with two challenges:

1. To expand the available sources - current trends include obtaining energy from
bio-fuels, hydrogen cells etc.

2. To utilize the available energy more efficiently.

It is important to understand the processes that govern transformation of the types


of energy mentioned, and the relationship between the associated work and heat, and
this is covered under thermodynamics.
Thermodynamics is that branch of science that deals with the interaction between
heat and work in a system. In engineering, thermodynamics is that tool that is nec-
essary for the understanding of energy, e.g. mechanical, electrical, chemical; and its
transformations, e.g. transformation of chemical energy into thermal energy. In ther-
modynamics, a wide range of energy systems can be analyzed using two primary laws:

• the First Law of Thermodynamics, and

• the Second Law of Thermodynamics

These will be discussed later. For a start, important terminologies that will be used in
this course are discussed next.

1.2 Terminologies in thermodynamics


1.2.1 Substance
As far as energy transformations are concerned, a substance is the matter that performs
the energy transformation. It may be a pure substance (homogeneous in nature, having
the same physical and chemical composition at all points in all conditions) e.g. oxygen,
or it may not be a pure substance e.g. air (nitrogen and oxygen condense at different
temperatures) [1].

3
1.2.2 System, boundary and surroundings
A system is a region of finite quantity of matter, or a space of fixed identity. A boundary
is the enclosing envelope for a system. It forms the interface between the system and
the surroundings - everything external to the system is known as the surroundings.
A system can be closed or open. If energy i.e., work and heat, can cross the
boundary but the boundary does not allow any matter exchange between the system
and the surroundings, then this is a closed system. For example, in Fig. 1.1, heat is

Figure 1.1: Power plant: closed system

added into the system in the steam generator (to vapourize the water into steam) and
lost from the system in the condenser (condensing spent steam back into water). On
the other hand, work is put into the system in the pump (in pumping the water into
the steam generator) while work is extracted from the system in the turbine (when
steam expands in the turbine to obtain mechanical work). However, there is no mass
transfer across the system boundary (dashed line).
An open system is one that allows matter, as well as work and heat to cross the
boundary. In Fig. 1.2, mass flows across the system boundary (dashed line) in the
form of air, fuel, cooling water, products of combustion. In some systems, there is no
work, heat or matter transfer across the boundary. Such a system is referred to as an
isolated system. If a system is thermally isolated from its surroundings (i.e., perfectly
insulated), then it is referred to as an adiabatic system, since no heat can cross the

4
Figure 1.2: Power plant: open system

system boundary [2].

1.2.3 Equilibrium
When there is no energy transfer between different parts of the mass of a fluid in
a system, and if the mass of the fluid is isolated from its surroundings e.g. if the
thermodynamic properties are constant, then the fluid is said to be in equilibrium.

• Thermal equilibrium: The system and its surroundings are at the same temper-
ature and there is no heat transfer across the boundary

• Mechanical equilibrium: There is uniform pressure or balance of forces between


the system and its surroundings

• Chemical equilibrium: There is uniform chemical composition between the system


and its surroundings.

1.2.4 Process, state and property


• A process is the transformation of a system from one state to another.

• The state of a substance completely describes how the substance exists. It com-
prises the temperature, pressure, density and other properties and by knowing
these properties, the state of the substance can be determined.

5
• The state of a system is the totality of its properties e.g. pressure, temperature,
density, specific energy etc.

• A property is an observable or calculable characteristic of a system e.g. density,


pressure, temperature, which depends not on how the system changes state, but
only on the final system state. There are intensive properties and extensive
properties

– An extensive property depends on the size or extent of the system e.g. mass
and volume.
– An intensive property is independent of the size of the system e.g. temper-
ature. An extensive property per unit mass e.g. specific volume (in m3 /kg)
is an intensive property.

1.2.5 Cycle and path


If the initial and final states of a process are the same, then this process is a cycle. A
cycle implies a path - this is a succession of system states. Consider the cycle in Fig.
1.1: The working fluid at state 1 is water. Mechanical work is put into the water in the
pump and the water is conveyed into the steam generator (or boiler) in state 2. In the
boiler, the water is vapourized into high-energy steam in state 3. This steam expands
in the turbine where its thermal energy is converted into mechanical work, and comes
out in state 4 as low-energy steam. This steam is then condensed in the condenser to
convert it back into water in state 1.

1.2.6 Work
The work done on or by a system, W, may be expressed as the product of force and the
distance moved in the direction of force. In Fig. 1.3, a piston is moved from the same
initial state (p1 ,v1 ) to the same end/final state (p2 ,v2 ) via different paths, a and b. If
the cross-sectional area of the piston is A and the distance the piston travels between
state 1 and state 2 is denoted dl, then,

δW = pAdl
= pdV
Z 2
W = pdV
1

Sign convention:
Work done by a system on the surroundings is positive.
Work done by the surroundings on the system is negative [2].

6
Figure 1.3: Work

1.2.7 Heat
Heat is defined as the energy crossing a system’s boundary because of a temperature
gradient between the system and its surroundings. Heat is represented by the symbol
Q.
The interaction at the system boundary is only due to temperature difference,
where heat is transferred from the hot system to the cold one. Heat is therefore a
transfer of thermal energy, which crosses the boundary of a system due to the difference
of temperature between the system and its surroundings.
Sign convention:
Heat transferred into a system is positive.
Heat transferred out of a system is negative.
Heat is a path function: it is a function of the method in which energy is transferred
between the system and the surroundings. A process in which there is no heat flow
across the boundary i.e. Q = 0, is termed adiabatic [3].

1.3 The Zeroth Law of Thermodynamics


Consider two bodies that are in contact with each other; one is hot, the other cold.
Heat will be lost by the hot body and gained by the cold one until a point of thermal
equilibrium is reached when the temperatures of both bodies will be equal. The zeroth
law of thermodynamics gives the scientific expression of bodies in the state of thermal
equilibrium and states that, If a system A is in thermal equilibrium with a system B
such that TA = TB , and if the system A is also in thermal equilibrium with a system C
such that TA = TC , then system B and system C must also be in thermal equilibrium,
i.e. TB = TC . It is therefore possible to compare the temperatures of two bodies by
use of a third body, typically a thermometer, without the two bodies being in physical

7
contact. The zeroth law of thermodynamics thus forms the basis for temperature
measurements [1, 3].

1.4 First Law of Thermodynamics


The first law of thermodynamics, also known as the conservation of energy principle,
provides a basis for studying the relationships among various forms of energy and
energy interactions. The law may be stated thus: Energy can neither be created
nor destroyed during a process, but it can change form.

1.4.1 Closed system - Non-flow energy equation


For a general closed system, the first law of thermodynamics may be stated as
1
Q − W = 4E = 4(U + mgz + mC 2 ), (1.1)
2
that is, the sum of work and heat interactions in a closed system is equal to the total
1
change in energy in that system. Often, 4mgz and 4 mC 2 , i.e. changes in potential
2
energy and kinetic energy respectively, are negligible in closed systems as compared to
4U (change in internal energy), and the first law then reduces to

Q − W = 4U (1.2)

This equation is also called the non-flow energy equation. The units of the energy terms
are Joules (J), or more commonly kiloJoules (kJ). The equation may be expressed in
terms of intensive properties as,

q − w = 4u, (1.3)

Q W U
where q = ,w= and u = (the extensive properties e.g. the internal energy U,
m m m
are converted into intensive properties e.g. the specific internal energy u, by dividing
by the mass m contained in the system). The units of the energy terms are now Joules
per kilogram (J/kg), or more commonly kiloJoules per kilogram (kJ/kg) [1, 2, 3].
In terms of differential changes of a system

δQ − δW = dU or
δq − δw = du.

The use of delta δ for q and w, and the normal differential d for u, is to distinguish
between an interaction across a system boundary (δ) and a change of a property within
the system (d ).

8
If the internal, potential and kinetic energies are restored to their initial values
by the heat and work interactions, then the right-hand side of Eq.(1.1) becomes zero.
Hence for a cycle, the total change in energy must be zero:

Q−W = 0 (1.4)

This is the statement of the first law of thermodynamics for a cycle.


EXAMPLES 1

1. In a certain steam plant operating in a closed cycle as in Fig. 1.1, the turbine
develops 1000 kW. The heat supplied to the steam in the boiler is 2800 kJ/kg,
the heat rejected by the system to the cooling water in the condenser is 2100
kJ/kg and the feed pump work required to pump the condensate back into the
boiler is 5 kW. Calculate the mass flow rate of steam round the cycle in kg/s.
[1.421 kg/s]

2. In the compression stroke of an internal-combustion engine, the heat rejected to


the cooling water is 45 kJ/kg and the work input is 90 kJ/kg. Calculate the
change in specific internal energy of the working fluid stating whether it is a gain
or a loss. [4u = 45 kJ/kg, gain]

HOMEWORK 1
In the cylinder of an air motor, the compressed air has a specific energy of 420 kJ/kg
at the beginning of the expansion and a specific internal energy of 200 kJ/kg after
expansion. Calculate the heat flow to or from the cylinder when the work done by the
air during the expansion is 100 kJ/kg. [q = -120 kJ/kg, i.e. 120 kJ/kg of heat flows
out]

1.4.2 Open system - Steady flow energy equation


In an open system (not necessarily steady), energy is transferred into and out of the
system not only by heat and work but also by the fluid that crosses into and subse-
quently out of the system, in the form of internal energy, gravitational potential energy,
kinetic energy and flow work. Some of the matter flowing through the system may also
be retained by the system, leading to accumulation.
The statement of the first law for such a general open system is: Energy entering
a system must equal the energy leaving the system plus any accumulation of energy
within the system.
In engineering, we often encounter problems concerning steady flow processes, i.e.
the rate at which the fluid flows through the machines or apparatus is constant and

9
there is no accumulation within the system [1, 2, 3]. The following conditions are
satisfied:

1. The mass rates of flow into and out of the system are equal and do not vary with
time

2. The energy of the fluid both at the entrance and exit does not change with time

3. The rates of heat and/or work transfer across the system boundary do not vary
with time

Figure 1.4: Open system

Therefore,

C12 C2
δ Q̇ − δ Ẇ + dṁ1 (u1 + p1 v1 + + gz1 ) = dṁ2 (u2 + p2 v2 + 2 + gz2 ) (1.5)
2 2
For steady flow, dṁ1 = dṁ2 = dṁ = mass flow rate (it is constant); and defining
enthalpy, h, as h = pv + u,

C12 C22
   
δ Q̇ − δ Ẇ + dṁ (h1 + + gz1 ) = dṁ (h2 + + gz2 ) (1.6)
2 2

This is the steady flow energy equation, which may also be conveniently written as,
 
1 2 2
Q̇ − Ẇ = ṁ (h2 − h1 ) + (C2 − C1 ) + g(z2 − z1 ) (1.7)
2

ṁ is the mass flow rate (in kg/s) and the units of the energy-flow components in the
equation are Joules per second (J/s) or Watts (W). In nearly all problems in applied
thermodynamics though, changes in height are negligible, thus the potential energy
terms can be omitted. The kinetic energy terms are also relatively small in many
applications and the whole equation then normally simplifies to,

Q̇ − Ẇ = ṁ(h2 − h1 ) (1.8)

10
EXAMPLE 2
In the turbine of a gas turbine unit, the gases flow through the turbine at 17 kg/s and
the power developed by the turbine is 14000 kW. The enthalpies of the gases at inlet
and outlet are 1200 kJ/kg and 360 kJ/kg respectively, and the velocities of the gases
at inlet and outlet are 60 m/s and 150 m/s respectively. Calculate the rate at which
heat is rejected from the turbine. Find also the area of the inlet pipe given that the
specific volume of the gases at inlet is 0.5 m3 /kg. [Q̇ = -119.3 kW; A1 = 0.142 m2 ]

Tutorial Problems 1

1. In an air compressor, the compression takes place at constant internal energy and
50 kJ of heat are rejected to the cooling water for every kg of air. Find the work
required for the compression stroke per kg of air. [50 kJ/kg]

2. A steam turbine receives a steam flow of 1.35 kg/s and delivers 500 kW. The
heat loss from the casing is negligible.

(a) Find the change of specific enthalpy across the turbine when the velocities
at entrance and exit and the difference in elevation at entrance and exit are
negligible. [370 kJ/kg]
(b) Find the change of specific enthalpy across the turbine when the velocity at
entrance is 60 m/s, the velocity at exit is 360 m/s, and the inlet pipe is 3 m
above the exhaust pipe. [433 kJ/kg]

3. A turbine operating under steady flow conditions receives steam at the following
state; pressure 13.8 bar; specific volume 0.143 m3 /kg; internal energy 2590 kJ/kg;
velocity 30 m/s. The state of the steam leaving the turbine is: pressure 0.35 bar,
specific volume 4.37 m3 /kg, internal energy 2360 kJ/kg, velocity 90 m/s. Heat is
lost to the surroundings at the rate of 0.25 kJ/s. If the rate of steam flow is 0.38
kg/s, what is the power developed by the turbine? [102.8 kW]

4. A nozzle is a device for increasing the velocity of a steadily flowing stream of fluid
(no work transfer is associated with a nozzle). At the inlet to a certain nozzle
the enthalpy of the fluid is 3025 kJ/kg and the velocity is 60 m/s. At the exit
from the nozzle the enthalpy is 2790 kJ/kg. The nozzle is horizontal and there
is negligible heat loss from it.

(a) Find the velocity at the nozzle exit. [688 m/s]

11
(b) If the inlet area is 0.1 m2 and the specific volume at inlet is 0.19 m3 /kg, find
the rate of flow of fluid. [31.6 kg/s]
(c) If the specific volume at the nozzle exit is 0.5 m3 /kg, find the exit area of
the nozzle. [0.0229 m2 ]

ASSIGNMENT 1 - Due on

1. Air flows steadily at the rate of 0.4 kg/s through an air compressor, entering at
6 m/s with a pressure of 1 bar and specific volume of 0.85 m3 /kg, and leaving at
4.5 m/s with a pressure of 6.9 bar and specific volume of 0.16 m3 /kg. The specific
internal energy of the air leaving is 88 kJ/kg greater than that of the air entering.
Cooling water in a jacket surrounding the cylinder absorbs heat from the air at
the rate of 59 kJ/s. Calculate the power required to drive the compressor, and
the inlet and outlet pipe cross-sectional areas.

2. A mass of gas with an internal energy of 1500 kJ is contained in a cylinder which


has perfect thermal insulation. The gas is allowed to expand behind a piston
until its internal energy is 1400 kJ. Calculate the work done by the gas. If the
expansion follows a law pV 2 = constant (polytropic process), and the initial
pressure and the volume of the gas are 28 bar and 0.06 m3 respectively, calculate
the final pressure and volume.

12
2.0 The Working Fluid
T he working fluid, or substance, is the matter contained within the boundaries of a
system. When two independent properties of the fluid are known, the thermodynamic
state of the fluid is defined.
In thermodynamic systems, the working fluid can be in liquid, vapour or gaseous
phase [1, 3].

2.1 Liquid, vapour and gas


Consider a p-v diagram for any liquid substance.
When the liquid is heated at constant pressure, there is a fixed temperature at which
vapour bubbles start forming - the liquid starts boiling. The higher the pressure, the
higher the boiling temperature.
A series of boiling points plotted on a p-v diagram appear as a sloping line as
in Fig. 2.1. Points P, Q and R are the boiling points at pressure pP , pQ and pR
respectively. When a liquid at boiling point is heated further at constant pressure, the

Figure 2.1: Liquid line

additional heat changes the phase of the liquid from liquid to vapour. The pressure
and temperature remain constant during the phase-change and the heat supplied is
called latent heat of vapourization. The higher the pressure, the smaller is the amount
of latent heat required.

13
For each constant pressure, there is a constant point of definite specific volume at
which vapourization is complete. For pressures pP , pQ and pR , these points are P0 , Q0
and R0 , and may be joined to form a line, Fig. 2.2. When Fig. 2.1 and Fig. 2.2 are

Figure 2.2: Vapour line

combined and extended to higher pressures, they form a continuous curve/loop with a
turning point called the critical point, point C on Fig. 2.3. The pressure at this point
is the critical pressure [3].

2.1.1 Saturation state


The saturation state is a state at which a change of phase may occur without change
of pressure or temperature e.g., P, Q, R, P0 , Q0 and R0 . The substance existing within
the loop e.g., at point S in Fig. 2.3, is referred to as wet vapour while the vapour
along the saturated vapour line is referred to as dry saturated, and no liquid is present
in the vapour in this state, i.e. along the vapourline R0 − Q0 − P 0 .
Lines of constant temperature, called isotherms, can be plotted on the p-v diagram
as in Fig. 2.4. These lines are horizontal in the wet vapour region. For each pressure,
there is a corresponding saturation temperature e.g., for pressure pQ , the saturation
temperature is T2 .
When a dry saturated vapour is heated at constant pressure, its temperature rises
and its vapour becomes superheated. The difference between the actual superheated
temperature and the saturation temperature at the pressure of the vapour is called the

14
Figure 2.3: Combined line

degree of superheat e.g., on Fig. 2.4, the vapour at point S is superheated at pressure
pQ and temperature T3 , and the degree of superheat is T3 − T2 , i.e. TS − TQ0 . For a wet

Figure 2.4: Isotherms

vapour, the temperature and pressure are not independent since they remain constant
for a range of values of the specific volume. The condition or quality of a wet vapour is
frequently defined by its dryness fraction. When the dryness fraction and the pressure,

15
or temperature, are known, then the state of the wet vapour is fully defined.
Dryness fraction, x = mass of dry vapour in 1kg of the mixture.

2.1.2 Perfect gas


At very high degrees of superheat, an isotherm on the p-v diagram tends to become a
hyperbola (obeying the law, pv = constant), e.g. the isotherm T6 on Fig. 2.4. When
pv
an isotherm follows a hyperbolic law, the equation of state =constant, is satisfied.
T
This is the equation of an idealized substance called a perfect gas.
The working fluid in practical engineering problems is either a substance approxi-
mating to a perfect gas e.g. air, or a substance that exists mainly as liquid and vapour
e.g. steam and refrigerant vapours such as ammonia and freon. For those substances
approximating to perfect gases, certain laws relating the properties can be assumed.
For the substances in the liquid and vapour phases, the properties are not related by
definite laws and values of the properties are determined empirically and tabulated in
a convenient form [1].

2.2 Vapour tables


These are the tables from which properties of substances that normally exist in the
vapour phase e.g. steam, can be read. They are normally called steam tables though
in essence they are Thermodynamic property tables. The commonly available tables
include properties for many substances other than water. These tables are normally
found in booklet form but they are also now available as computer programs [4].
For steam:
The saturation pressures and corresponding saturated temperatures are tabulated in
parallel columns for pressures ranging from 0.006112 bar to the critical pressure of
221.2 bar.
The specific volume, specific internal energy, specific enthalpy and specific entropy, are
also tabulated for the saturated vapour at each pressure and corresponding saturation
temperature. Suffix f is used to denote values on the saturated liquid line while suffix
g is used to denote the dry saturated state. To denote the change of property between
the saturated liquid line and the saturated vapour line, a suffix fg is used. When the
property is specific enthalpy h, then hf g is the latent heat, where hf g = hg − hf [1, 3].

16
2.2.1 Wet vapour
volume of liquid + volume of dry vapour
1. Specific volume, v is given by v = .
total mass of wet vapour
For 1kg of wet vapour with dryness fraction of x, there are x kg of dry vapour and
(1-x )kg of liquid. Hence, v = (1 − x)vf + xvg . But normally, the volume of liquid
is usually negligibly small compared to the volume of dry saturated vapour, i.e.,
vf  vg , hence

v = xvg (2.1)

2. The enthalpy of a wet vapour is given by the sum of the enthalpy of the liquid
plus the enthalpy of the dry vapour

h = (1 − x)hf + xhg
= hf + x(hg − hf )
= hf + xhf g (2.2)

3. The internal energy of a wet vapour is given by the internal energy of the liquid
plus that of the dry vapour

u = (1 − x)uf + xug
= uf + x(ug − uf ) (2.3)

Note that uf g is not tabulated in the steam tables unlike the case for enthalpy.

EXAMPLE 3
Determine the dryness fraction, specific volume and specific internal energy of steam
at 7 bar and enthalpy 2600 kJ/kg. [x=0.921; v=0.2515 m3 /kg; u=2420 kJ/kg]

2.2.2 Superheated vapour


In the superheat region, temperature and pressure are independent, and both must be
given to define the state of the vapour.
EXAMPLES 4

• At 2 bar, the saturation temperature is 120.20 C, therefore, steam at 2 bar and


2000 C is superheated; and the degree of superheat is 2000 C - 120.20 C = 79.80 K.

• At 20 bar and 4000 C, the specific volume is 0.1511m3 /kg and the enthalpy is
3248kJ/kg.

Note that above 70 bar, values of the internal energy are not tabulated and the equation:
h = u + pv, has to be used, e.g.

17
• For steam at 80 bar and 4000 C, the enthalpy h = 3139kJ/kg, specific volume, v
is 3.428 x 10−2 m3 /kg and so,

u = h − pv
80x105 x 0.03428
= 3139 −
103
= 2864.8 kJ/kg

Tutorial Problems 2

1. Steam at 150 bar has a specific enthalpy of 3309 kJ/kg, find the temperature,
specific volume and the specific internal energy. [5000 C; 0.02078 m3 /kg; 2997.3
kJ/kg]

2. 0.05 kg of steam at 15 bar is contained in a rigid vessel of volume 0.0076 m3 .


What is the temperature of the steam? If the vessel is cooled until the pressure
in the vessel is 11 bar, calculate the dryness fraction of the steam and the total
heat rejected. [2500 C; 0.857; 18.5 kJ]

2.2.3 Interpolation
For properties which are not tabulated exactly in the tables, it is necessary to interpo-
late between the values tabulated e.g.

• Find the temperature, specific volume, internal energy and enthalpy of dry sat-
urated steam at 9.8 bar.
Solution:
9.8 bar falls between 9 bar and 10 bar, hence
9.8 − 9
T9.8bar = T9bar + (T10bar − T9bar )
10 − 9
9.8 − 9
= 175.4 + (179.9 − 175.4)
10 − 9
= 1790 C

Similarly,
9.8 − 9
hg,9.8bar = hg,9bar + (hg,10bar − hg,9bar )
10 − 9
= 2774 + 0.8(2778 − 2774)
= 2777.2 kJ/kg

18
2.3 Thermodynamic charts/diagrams
Tables of thermodynamic properties provide accurate data for various substances. But
the data from, say, steam tables, are equilibrium data. Charts plotted from these data
represent only the equilibrium states. The path of a process which is not an equilibrium
path cannot be drawn on these charts. Such processes are best portrayed in charts such
as temperature-entropy (T-s) charts and enthalpy-entropy (h-s) charts [5].

2.3.1 T-s diagram


Figure 2.5 shows a T-s chart for steam. It shows the liquid phase and the vapour
phase.

Figure 2.5: Outline of a T-s diagram for steam

Features

19
• Wet region (below saturation curve)

– Lines of constant temperature and constant pressure are horizontal


– Lines of constant dryness and constant volume are also shown

• Superheat region (above saturation curve)

– Lines of constant pressure start at the saturation curve, then rise steeply to
almost vertical
– Lines of constant enthalpy are almost horizontal away from the saturation
curve. Near the saturation curve (especially near the critical point), there is
a marked change of curvature of the constant enthalpy lines, which approach
the vertical as the critical pressure is approached.

Though the T-s chart is useful in portraying processes, it is not as useful as the h-s
diagram (Mollier chart) [5].

2.3.2 Mollier Chart


The Mollier chart is a plot on h-s coordinates of the thermodynamic properties of a
substance. Figure 2.6 shows the outline of a Mollier chart for steam [5]. The Mollier
chart is particularly suited to:
-obtaining properties
-describing flow

EXAMPLES 5

1. Determine the enthalpy and entropy of steam at 2 MPa and 2400 C, using the
Mollier chart. Compare your values with those obtained using steam tables.
[h=2880 kJ/kg, s=6.5 kJ/kgK; (h=2876.1 kJ/kg, s=6.492 kJ/kgK)]

2. Determine the enthalpy of saturated steam at 300 C using the tabulated properties
of pressure, specific volume and internal energy. Compare the results with the
tabulated value of hg . [p = 4.246 kP a, vg = 32.894 m3 /kg, ug = 2416.6 kJ/kg,
h=u+pv=2556.27 kJ/kg; tabulated value of hg = 2555.7 kJ/kg]

3. Repeat problem 2 above from the Mollier chart

Tutorial Problems 3

1. Why is high moisture content in steam detrimental to steam turbines?

20
Figure 2.6: Outline of Mollier chart for steam

2. What is the relationship between steam moisture content and quality?

3. Wet steam at 1.0 MPa is found to have a quality of 85%.


(a) Use Steam Tables to determine its entropy and specific volume.
(b) Use the Mollier chart to determine the enthalpy.
(c) From the results above, calculate the internal energy.

21
Figure 2.7: Mollier chart

22
3.0 The Second Law Of Thermodynamics
According to the first law of thermodynamics, when a closed system undergoes a
complete cycle, then the net heat supplied to the system is equal to the net work done
by the system, Q − W = 0. This is based on the conservation of energy and follows
from the observation of natural events. The second law of thermodynamics, which is
also a natural law, indicates that although the net heat supplied in a cycle is equal to
the net work done, the gross heat supplied must be greater than the net work done,
meaning that some heat must always be rejected by the system. We will consider a
heat engine [2, 3].

3.1 The heat engine


The heat engine is a system operating in a complete cycle and which converts heat
input into work output. The piston-cylinder heat engine is the most common heat
engine.
The system for the piston-cylinder engine will be the air/gas contained therein.
We will assume the gas is ideal (or perfect) such that it obeys the equation of state,

pv = RT (3.1)
Ru
R is the specific gas constant and is calculated, R = where Ru is the universal gas
M
3
constant (8.314 x 10 J/kmol K) and M is the molecular weight of the gas (which is 29
8.314 x 103
kg/kmol for air). For air, R = = 2.87 x 102 J/kgK.
29

Figure 3.1: Heat engine. The gas inside the cylinder is the system

We will consider the system to be closed (no mass transfer across the boundary).
Let a hot plate supply heat to the gas in such a way that all the heat goes into the gas

23
and none is absorbed by the cylinder wall - refer to Fig. 3.1 (a). The system undergoes
energy changes; the potential energy and kinetic energy changes are however assumed
small, compared to the internal energy change [2].
When the gas is heated, it will expand and push against the piston. Let the piston
initially be restrained so that it does not rise. This causes the pressure of the gas to
increase between 1 and 2, Fig. 3.2. As the pressure reaches point 2, the restraint
is removed and the piston will rise thus lifting the weight mg and doing work. As
the piston does work, the process could proceed in an infinite number of ways e.g.
at constant pressure, 2-3a, or it could be polytropic, pv n = constant, where n is a
polytropic index (process 2-3b). The particular path taken affects the efficiency of the
engine. Let us assume the path taken is 2-3b. At point 3b, the weight is taken off;

Figure 3.2: pv diagram for the heat engine

work has already been done in lifting the weight.


The piston will now be returned to its original state in two processes because the
pressure acting on it may be high: First, some heat is taken away from the gas through
a cold plate. This is at constant volume, process 3b-4 (corresponding to Fig. 3.1 c).
When the pressure has come sufficiently down, at point 4, the piston can now move
down against the gas in the process 4-1 thus taking the system back to its initial state

24
[2].
There is also an infinite number of possible paths for process 4-1, just like in the
process 2-3.

• The work done by the system during process 2-3b is given as,
Z 3b
W2−3b = m pdv. (3.2)
2

Note that because the system is closed, m is constant.

• Work done on the system in the process 4-1 is,


Z 1
W4−1 = m pdv (3.3)
4

In both cases, the relationship between p and v must be known for the respective
equations to be integrated. The essential point to note is that the work done is
proportional to the area under the curve on the p-v plot.

• The heat input into the system, is represented by process 1-2 and is calculated,

Q1−2 = mcv (T2 − T1 ) (3.4)

cv is the specific heat capacity of the gas at constant volume (note that the
heating occurs at constant volume).

• The thermal efficiency is defined as the ratio of the net work output to the heat
input into the system. In this case,
Wnet W2−3b − W4−1
ηth = = (3.5)
Qin Q1−2
A basic issue in heat engine design has always been: What is the best path the
processes should take, so as to produce the most efficient engine? This is a central
concern of the second law of thermodynamics.

It is not possible to have an efficiency of 100% since some of the heat input is always
dissipated in such processes as overcoming friction. Nevertheless, engineers have always
tried to modify designs e.g. changing pressure ratios, varying the load on the piston
etc, to improve efficiency. Two of the most common cycles are:

1. The Otto cycle: this describes the typical automobile internal combustion (spark-
ignition) engine

2. The Diesel cycle: typical for large vehicles such as trucks and buses that use the
compression-ignition engines

25
3.2 Reversibility
The heat engine already described does not have a thermal efficiency of 100%. For
instance, some of the heat input is used to overcome friction. Such factors lead to
process irreversibilities. One way of simulating reversibility is by pulling the piston a
small amount to the right so slowly that we do not cause any large-scale motion of the
gas, Fig. 3.3 (a). The process can then be reversed, i.e., the piston may be pushed
back to its original position so that everything returns to the same state it started
from. Such an ideal work transfer is regarded a reversible work interaction, where all
properties are restored to their initial values after the pull-push motion is complete. In
this process, friction is assumed absent between the cylinder and the piston (friction
causes a process to be irreversible).

Figure 3.3: (a)Reversible and (b)Irreversible work transfer

In reality, all processes are irreversible, but the concept of reversibility is central to
thermodynamic analysis. Reversibility implies that everything is infinitesimally close
to equilibrium at all times throughout the process.
As far as heat interaction is concerned, if the heat interaction occurs at virtually
constant temperature, with only an infinitesimal difference to provide the heat transfer
in the required direction, the heat interaction is reversible. In Fig. 3.4 (a), if we heat
the object a little and cool it a little, the heat will flow first in one direction and then
in the other direction in a reversible way. In other words, to attain reversible heat
transfer, we decrease the temperature difference until in the limit, it is almost zero
(infinitesimally small difference). Heat transfer across a finite temperature difference
as in Fig. 3.4 (b) is irreversible.
In summary, the concept of reversibility means that the system and its surroundings
are always infinitesimally close to equilibrium [2]. A reversible process then is any
process performed so that the system and all its surroundings can be restored to their
initial states by performing the process in reverse [5].

3.3 Deriving the second law of thermodynamics


The essence of a heat engine is that it has a heat intake from a high temperature
source, and a heat rejection to a low-temperature sink. We will call these the high and

26
Figure 3.4: (a)Reversible and (b)Irreversible heat transfer

low temperature reservoirs respectively [2].

Figure 3.5: (a)Schematic diagram of a heat engine. (b)A reversible heat engine (c)A
heat pump - a heat engine working as a heat pump or refrigerator

Consider Fig. 3.5 (a) showing heat input from a high temperature reservoir, heat
rejection to a low temperature reservoir and net work done on surroundings. If the

27
cycle were reversible, all the arrows of Fig. 3.5 (a) can be reversed, as shown in Fig.
3.5 (b). A heat engine working in reverse is called a heat pump, requiring work input
to transfer heat from a low temperature to a high temperature reservoir. If a heat
pump is irreversible, it would be represented by Fig. 3.5 (c).

Figure 3.6: (a)A reversible heat engine (ER ) and an irreversible heat engine (E) both
operating at the same temperature difference, and both extracting the same amount of
heat from TH (b)The reversible engine is now reversed (c)The high temperature source
TH is unnecessary, and net work is produced by extracting heat from a low-temperature
reservoir only.

Now consider the following steps:

• Let us place a reversible engine ER and an irreversible one E side by side, to


operate between the same hot and cold plates (TH and TC respectively) as shown
in Fig. 3.6 (a). The engines do work WR and W respectively, while extracting
the same amount of heat, QH .

• Let us reverse ER (Fig. 3.6 (b)) and use symbol ∃R for it. The engine now
transfers heat QH to TH while an amount of work WR is done on it. Since the
heat QH rejected by ∃R to TH is the same amount extracted by E from TH , then
the reservoir TH becomes redundant. Again, if we assume W is greater than WR ,

28
then E can run ∃R and still have some work W − WR left over to do external
work. This situation is represented by Fig. 3.6 (c).

Such a system, though not violating the first law of thermodynamics, still violates
natural expectations: It does not require a temperature difference and so requires no
fuel to provide a high-temperature source. If it could exist, all energy and environmen-
tal problems would vanish! The first law is obeyed thus,

Q − W = 0, i.e., (QR − Q) − (W − WR ) = 0 (3.6)

For each of the engines, we would have:

• For the reversible engine ∃R ,

(QR − QH ) + WR = 0 (3.7)

• For the irreversible engine E,

(QH − Q) − W = 0 (3.8)

Adding Eqs. (3.7) and (3.8) gives Eq. (3.6). From Eq. (3.6), QR − Q = W − WR ,
and because we have assumed W > WR , then QR > Q.
The flaw is in assuming W > WR , i.e., assuming that an irreversible engine can
produce more work than the reversible engine operating between the same hot and
cold reservoirs (Fig. 3.6 (a)). If we assume W < WR , then it is fine, but it would
similarly mean that QR < Q. This would be a useless friction machine, that simply
converts work into heat.
The second law must then be stated: It is impossible for any system to
operate in a thermodynamic cycle and do net work on its surroundings,
while exchanging heat with a single reservoir [2].
There are many statements and corollaries of the second law that can be found in
thermodynamics literature. For example, two other statements are:

• Clausius statement: It is impossible to construct a device which operates in


a cycle and whose sole effect is to transfer heat from a cooler body to a hotter
body [5]

• Kelvin-Planck statement: It is impossible to construct a device which operates


in a cycle and produces no other effect than the production of work and exchange
of heat with a single reservoir [5]

29
3.4 Entropy
The second law of thermodynamics gives rise to a property called entropy. Entropy is
a thermodynamic property (does not depend on the path taken). It is denoted by S
while specific entropy is denoted by s. The defining equation for entropy is
dQ
ds = , for a reversible process (3.9)
T
This applies to all working substances [3].
With regard to entropy, another statement of the second law is: In any physical
process, the total entropy never decreases. All natural processes are regarded as
irreversible [5] and whenever energy is used, it is degraded [4]. In an irreversible process,
entropy increases and the gain in entropy is regarded as a measure of irreversibility.
For reversible adiabatic processes, the entropy remains constant. Such processes are
called isentropic.
In other applications, entropy is regarded as a measure of randomness, e.g., in
statistical mechanics, the Universe is considered to have evolved from a more ordered
to a less ordered state and so entropy is deemed to have increased.

3.5 The Carnot cycle


It can be shown from the second law that no heat engine can be more efficient than
a reversible heat engine working between the same temperature limits. Sadi Carnot,
a French engineer, showed that the most efficient possible cycle is one in which all
the heat supplied is supplied at a fixed upper temperature, and all the heat that is
rejected is rejected at a lower fixed temperature. The complete cycle consists of two
isothermal processes and two adiabatic processes. Since all the processes are reversible,
the adiabatic processes in the cycle are termed isentropic (reversible and adiabatic).
The processes are,
1-2: isothermal heat supply
2-3: isentropic expansion
3-4: isothermal heat rejection
4-1: isentropic compression
The Carnot engine and p-v cycle are shown in Fig. 3.7, while the T-s diagram is
shown on Fig. 3.8.
The thermal efficiency of the cycle is,
Wnet Wnet
ηth = = (3.10)
Qin QH

30
Figure 3.7: The Carnot engine

Figure 3.8: Carnot T-s cycle

From the first law of thermodynamics for a cycle, Q - W = 0, i.e.

(QH − QC ) − (Wout − Win ) = 0


QH − QC − Wnet = 0
QH − QC = Wnet (= Qnet ) (3.11)

Therefore,
QH − QC Qnet QC
ηth = = =1− , (3.12)
QH QH QH
but
QC TC (s3 − s4 )
= but (s2 − s1 ) = (s3 − s4 ), (3.13)
QH TH (s2 − s1 )

31
therefore,
TC
ηth = 1 − (3.14)
TH
The Carnot cycle efficiency is the highest that can be achieved between any two fixed
temperatures. All others must be less than it.
EXAMPLE 6
What is the highest possible theoretical efficiency of a heat engine operating with a
hot reservoir of furnace gases at 20000 C when the cooling water available is at 100 C?
[87.54%]
Tutorial Problem 4
A Carnot engine operates between a source temperature of 7000 C and a sink temper-
ature of 200 C. Assuming that the engine will have a net output of 65 hp, determine
the thermal efficiency of the engine, the heat supplied and the heat rejected. [69.9%;
69.37 kJ/s; 20.88 kJ/s]
To increase the Carnot efficiency, the temperature difference TH − TC can be in-
creased but there are limitations e.g., for a fixed TC , TH can only be raised to as high
a value as the metallurgical capabilities of the material components of the engine.
Again, in practice, it is difficult to devise a system which can receive heat at a
constant temperature, and reject heat at a constant temperature. A wet vapour is
the only working substance which can do this conveniently, since for a wet vapour the
pressure and temperature remain constant as the latent heat is supplied or rejected.
Such a cycle is shown in Fig. 3.9 [1].

Figure 3.9: A T-s diagram for a Carnot cycle using steam

32
3.6 Absolute temperature scale
It is possible to establish a temperature scale that is independent of the working fluid.
Recall from the efficiency of a Carnot cycle (Eq. (3.12)) that

QH − QC Qnet
ηth = =
QH QH
The efficiency of an engine operating on the Carnot cycle depends only on the temper-
atures of the hot and cold reservoirs. Let us denote temperature on an arbitrary scale
by X such that

ηth = φ(XH , XC ) (3.15)

where φ is a function while XH and XC are the temperatures of the hot and cold
reservoirs respectively.
Combining Eqs. (3.12) and (3.15),

Qnet
= φ(XH , XC ) (3.16)
QH
The function φ cannot be determined analytically, for it is entirely arbitrary and many
temperature functions can satisfy it [5]. Kelvin proposed that the temperature function
may be chosen thus,
Qnet XC
1− = = 1 − ηth
QH XH
XC
∴ ηth = 1 − (3.17)
XH
From Eq. (3.14), we have

TC
ηth = 1 −
TH
or,
Qnet TC
1− = (3.18)
QH TH

Comparing Eq. (3.17) and (3.18), it can be seen that the temperature X is equivalent
to the temperature T . Thus, by suitably choosing the function φ, the ideal temperature
scale is made equivalent to the temperature scale based on the perfect gas. Such a scale
is the absolute thermodynamic temperature scale and does not depend on the working
substance [3, 5].

33
4.0 Vapour Power Cycles
4.1 The Rankine cycle
T he Carnot cycle is the most efficient cycle for given temperatures of the source and
sink, and applies to both gases and vapours. However,the Carnot cycle is not normally
used in steam plants. One reason is that it has a low work ratio,
net work
work ratio =
gross work
For the steam cycle of Fig. 3.9, per unit mass of substance,

net work = wout − win


= (h2 − h3 ) − (h1 − h4 )
gross work = wout = (h2 − h3 )

therefore,
(h2 − h3 ) − (h1 − h4 ) h1 − h4
work ratio = = 1−
(h2 − h3 ) h2 − h3
compression work
= 1− (4.1)
expansion work
The other reason for not using the Carnot cycle in the steam plants is because, at state
4, the steam is wet. It is difficult to stop condensation at point 4 and then compress
it to state 1, Fig. 3.9. It is more convenient to allow the condensation process to
proceed to completion, i.e. to the saturated liquid line, as in Fig. 4.1.
The working fluid is now water at state 4, Fig. 4.1, and is conveniently pumped
to the boiler at state 1. The pump is now smaller and compression work input, h1 − h4 ,
is now reduced as compared to that of Fig. 3.9. When compression work is reduced,
then from Eq. (4.1) the work ratio is increased. This modified cycle represented by
Fig. 4.1 is the Rankine cycle.
The disadvantage is that more heat input is now required (h2 − h1 ) as opposed to
that of the Carnot cycle in Fig. 3.9. From Eq. (3.10), the thermal efficiency of the
cycle is thus reduced.
For the Rankine cycle, per unit mass,

• Heat input in the boiler, qin = h2 − h1

• Isentropic steam expansion in the turbine does work, wout = h2 − h3

• Heat rejection from the steam in the condenser, qout = h3 − h4

34
Figure 4.1: T-s diagram for the Rankine cycle

• Isentropic compression in the pump, win = h1 − h4 ≈ vf,p4 (p1 − p4 ), NB:p1 =


p2 and p3 = p4

4.2 Rankine cycle with superheat


Between states 2 and 3 of Fig. 4.1, the turbine has to handle low-quality steam,
which has a lot of moisture content. The moisture leads to erosion and corrosion of
the turbine blades.
But since the heating occurs at constant pressure, then one way of reducing the
moisture that the turbine has to handle is by superheating the steam at state 2, as
shown in Fig. 4.2. A typical value of the temperature of the steam at state 2 is 500-
6000 C. Metallurgical limitations prevent attainment of higher values. A wide range of
pressures can be used though.
EXAMPLES 7

1. A steam power plant operates on the ideal Rankine cycle with superheat. The
steam enters the turbine at 7 Mpa and 5500 C. It discharges to the condenser at
20 kPa. Determine the cycle thermal efficiency [37.7%]

2. Steam is supplied dry saturated at 40 bar to a turbine, and the condenser pressure
is 0.035 bar. If the plant operates on the ideal Rankine cycle, calculate the work
output from the turbine, neglecting the feed pump work [986 kJ/kg].
Without neglecting feed pump work, determine

35
Figure 4.2: Rankine cycle with superheat

(a) the heat supplied [2685 kJ/kg], and


(b) the Rankine efficiency [36.6%]

Tutorial Problems 5

1. Why is the Carnot cycle not used as the ideal model for steam power plants?

2. What are the processes that make up the ideal Rankine cycle?

3. Explain the effect of lowering the condenser pressure in an ideal Rankine cycle
on; turbine work, heat added, and thermal efficiency.

4. A Carnot cycle uses steam as the working substance and operates between pres-
sures of 7 Mpa and 7 kPa. Determine

(a) the thermal efficiency [44.2%]


(b) the turbine work [968.4 kJ/kg]
(c) the compressor work [303.7 kJ/kg]
(d) the work ratio [0.686]

5. A steam power plant operates between a boiler pressure of 42 bar and a condenser
pressure of 0.035 bar. Assuming ideal conditions, calculate the cycle efficiency
when the steam enters the turbine superheated at 5000 C. Neglect the feedpump
work [39.9%]

36
4.3 Isentropic efficiencies
Irreversibilities are associated with each of the components of the Rankine cycle. Fluid
friction and heat loss to the surroundings are the most common causes of irreversibility.

4.3.1 Turbine
In the turbine, the ideal expansion process is isentropic, but as the steam flows through
the turbine blading, fluid friction occurs, increasing the steam’s entropy. Fluid flow
irreversibilities significantly reduce the useful turbine work output. The turbine inter-
nal or isentropic efficiency is what accounts for the irreversibility losses. The turbine
isentropic efficiency is thus a determination of how well the available energy is used.
The value is determined experimentally by the turbine manufacturer and once known,
it may be used to compute the actual work the steam does in the turbine. Denoting
the turbine efficiency as ηT , then referring to Fig. 4.3,
h2 − h30
ηT = (4.2)
h2 − h3
Process 2 − 30 is the actual steam expansion in the turbine while process 2 − 3 is the
isentropic expansion.

Figure 4.3: Isentropic efficiencies

37
4.3.2 Pump
In the pump, frictional effects mean more work is required than the ideal case to raise
the water’s pressure to a higher value. These frictional effects in fluid flow through the
pump and between the pump’s impeller and the water increase the entropy. Heat loss
in the pump may be considered negligible. Let the pump isentropic efficiency be ηC .
With reference to Fig. 4.3,
h1 − h4
ηC = (4.3)
h10 − h4
In Rankine cycles, the pump work is much smaller than the turbine work, so the net
effect of pump inefficiency on cycle efficiency is small [3].
Tutorial Problems 6

1. The maximum steam temperature is 5600 C and the lowest cycle temperature is
300 C.

(a) What is the thermal efficiency of an ideal Rankine cycle operating with a
maximum pressure of 3.5 MPa? [39.6%]
(b) For the same temperatures and pressure above, recalculate the thermal ef-
ficiency for a Rankine cycle with a turbine efficiency of 80% [31.65%]

2. Discuss the Clausius inequality in the context of the second law of thermody-
namics

3. (a) If the condenser pressure in a Rankine cycle is 1 psia and the maximum
pressure in the cycle is 600 psia, calculate the efficiency of the ideal cycle if
the steam at 600 psia is dry saturated vapour. Use the Mollier chart and
neglect pump work. [34.9%]
(b) If the vapour above has 2000 F superheat, calculate the efficiency of the ideal
Rankine cycle.

4. In a Rankine cycle, the lower operating and condenser pressure is 3 bar. What
is the specific work output for an isentropic efficiency of expansion of 80%, if the
steam is generated at 30 bar and 3250 C? [379 kJ/kg]

5. In a steam power plant, steam leaves the boiler at a pressure of 16 MPa and a
temperature of 10000 F. The steam passes through a chest of throttle valves where
the pressure drops by 3%. The steam then expands in a high pressure turbine
at an isentropic efficiency of 80%, to 2 bar. Use steam tables to determine the
enthalpy of the steam at the exit from the turbine. [2947 kJ/kg]

38
4.4 Steam Turbines in Cogeneration
Cogeneration is defined as the production of more than one useful form of energy from
the same energy source, e.g., electric power and process heat obtained from steam [6].
Some industries that normally rely quite heavily on process heat include chemical,
pulp and paper, oil production and refining, steel making, food processing, and textile
industries. This heat is usually supplied by steam at 5 - 7 atm and 150 - 2000 C.
Ordinarily, this steam is produced by burning coal, natural gas etc in a furnace. The
temperatures in furnaces is normally very high (around 14000 C), meaning that the
energy in the furnace is high quality energy. Yet, this high-quality is only utilized to
produce steam at around 2000 C or less. It thus makes more economical and engineering
sense to produce the process steam using low-quality energy, thus the cogeneration
plant.
A typical example of a cogeneration plant is a steam-turbine cogeneration plant,
shown schematically in Fig. 4.4 with typical operating conditions. The plant produces
net work output Ẇnet via the turbine, and then the exhaust energy from the turbine is
used to generate the process heat Q̇p . The pump work is usually very small and may
be safely neglected. Notice also that the plant does not have a condenser; no heat is
rejected from the plant as waste heat, i.e., all the energy transferred to the steam in
the boiler is utilized as either process heat or electric power. We may then define a
utilization factor u for the cogeneration plant as
net work output + process heat delivered Ẇnet + Q̇p
u = = , (4.4)
total heat input Q̇in
or
Q̇out
u = 1 − , (4.5)
Q̇in
where besides the heat lost in the condenser, Q̇out includes all the undesirable heat
losses, even from the piping and other system components (these losses are normally
neglected).
The plant represented above is ideal though. In a typical practical steam co-
generation plant (see Fig. 4.5), some steam is extracted from the turbine at some
predetermined pressure, say p6 and used for process heating. The rest of the steam
expands to the condenser pressure, p7 and is then cooled at constant pressure. The
heat rejected from the condenser represents the waste heat for the cycle.
NOTES

• In times of high demand for process heat, all the steam leaving the turbine is
routed towards process-heating and none goes to the condenser (ṁ7 = 0).

39
Figure 4.4: An ideal steam-turbine cogeneration plant

• Maximum process heating is realized when all the steam leaving the boiler passes
through the expansion (throttle) valve. In that case, ṁ5 = ṁ4 and no power is
produced.

• When there is no demand for process heat, all the steam passes through the
turbine and the condenser such that ṁ5 = ṁ6 = 0. In that case, the cogeneration
plant operates as an ordinary steam power plant.

40
Figure 4.5: A practical cogeneration plant

• Pertinent equations

Q̇in = ṁ3 (h4 − h3 ), (4.6)


Q̇out = ṁ7 (h7 − h1 ), (4.7)
Q̇p = ṁ5 h5 + ṁ6 h6 − ṁ8 h8 , (4.8)
Ẇturb = (ṁ4 − ṁ5 )(h4 − h6 ) + ṁ7 (h6 − h7 ). (4.9)

ASSIGNMENT 2 - Due on
Consider the steam cogeneration plant shown in Fig. 4.6. Steam enters the turbine
0
at 7 MPa and 500 C. Some steam is extracted from the turbine at 500 kPa for process
heating. The remaining steam continues to expand to 5 kPa. The exhaust steam from

41
the turbine is subsequently condensed at constant pressure and pumped to the boiler
pressure of 7 MPa. At times of high demand for process heat, some heat leaving the
boiler is throttled to 500 kPa and then routed to the process heater. The extraction
fractions are adjusted so that the steam leaves the process heater as a saturated liquid
at 500 kPa and subsequently pumped to 7 MPa. The mass flow of steam through
the boiler is 15 kg/s. Neglecting the changes in potential and kinetic energies, and
disregarding the pressure drops and heat losses in the piping, and assuming the turbine
and pump work to be isentropic, determine:

1. The specific enthalpy of the steam at each of the states 1 through to 10

2. The maximum rate at which process heat can be supplied

3. The power produced and the utilization factor when no process heat is supplied

Figure 4.6: Schematic diagram for the cogeneration plant

42
5.0 Heat Transfer
Heat transfer is defined as energy in transit due to a temperature difference, i.e.,
whenever there exists a temperature difference in a medium, or between media, heat
transfer must occur.
The different types of heat transfer are called modes and there are three modes:
conduction, convection and radiation.

5.1 Conduction
This is the mechanism of internal energy exchange from one body to another, or from
one part of a body to another, by exchange of kinetic energy of motion of the molecules
by direct contact, or by drift of the free electrons in the case of heat conduction in
metals.
The flow of energy is from the higher energy molecules to the lower energy ones,
i.e., from a high temperature region to a low temperature region e.g. the exposed end of
a metal spoon immersed in a cup of hot coffee eventually warms up due to conduction
of heat energy through the spoon. Other examples?
Conduction occurs in both solids and fluids.
Heat transferred by conduction may be quantified using a rate equation called
Fourier’s law. This equation is expressed as:
dT
qx0 = −k (5.1)
dx
For a one-dimensional case, the heat flux, qx0 (in W/m2 ) is the heat transfer rate in the
x-direction per unit area (perpendicular to the direction of transfer), and is proportional
dT
to the temperature gradient .
dx
The proportionality constant k is a transport property called thermal conductivity
(in W/m K) and is a characteristic of the material (independent of the geometry). If
the temperature distribution is linear as in Fig. 5.1,
dT T2 − T1
= ,therefore, (5.2)
dx L
T2 − T1
qx0 = −k (5.3)
L
If the surface area is A, then the heat rate by conduction, qx (in W), is calculated thus,

T2 − T1
qx = −kA ,i.e., (5.4)
L
qx = qx0 .A (5.5)

43
Figure 5.1: One-dimensional conduction

EXAMPLE 8
The wall of an industrial furnace is constructed from 0.15 m-thick fire-clay brick hav-
ing a thermal conductivity of 1.7 W/m K. Measurements made during steady-state
operation reveal temperatures of 1400K and 1150K at the inner and outer surfaces
respectively. What is the rate of heat loss through a wall that is 0.5 m by 3 m on one
side? [4.25 kW]

5.1.1 Multilayer wall


Consider a plane wall composed of layers of material having different thicknesses and
thermal conductivities, There exists an analogy between diffusion of heat and electrical
charge. Just as an electrical resistance is associated with the conduction of electricity,
a thermal resistance may be associated with conduction of heat.
Considering Eq. 5.4 for example, the thermal resistance is given thus,
T1 − T2 L
Rconduction = = (5.6)
qx kA
Remember Ohm’s law gives an electrical resistance thus,
E1 − E2 L
Relec = = , (5.7)
I σA
where σ denotes the electrical conductivity, which is the inverse of resistivity. From
LA LB LC
Fig. 5.2, RA = , RB = and RC = .
kA A kB A kC A

44
Figure 5.2: Multilayer plane wall

It can be shown that:


T1 − T4
qx0 = , (5.8)
ΣR
where ΣR = RA + RB + RC .
ASSIGNMENT 3 - Due on
A house wall consists of an outer layer of common brick 10 cm thick (k = 0.69 W/m0 C),
followed by a 1.25 cm layer of Celotex sheathing (k = 0.048 W/m0 C). A 1.25 cm layer
of sheetrock (k = 0.744 W/m0 C) forms the inner surface and is separated from the

45
sheathing by 10 cm of air space. The air space has a unit conductance of 6.25 W/m0 C.
The outside brick temperature is 50 C; the inner wall surface is maintained at 200 C.
What is the rate of heat loss, per unit area of the wall?
What is the temperature at a point midway through the Celotex layer?

5.1.2 Hollow cylinder


Hollow cylinders are of great importance in engineering. If the inner surface of radius

Figure 5.3: Hollow cylinder

r1 and outer surface of radius r2 are maintained at uniform temperatures of T1 and T2


respectively, then for a sufficiently long cylinder, the temperature is independent of the
z co-ordinate. The problem is reduced to a one-dimensional case (in the steady state)
with the radial distance r as the co-ordinate. Fourier law for such a cylinder is:
dT dT
qr = −kA = −k(2πrL) (5.9)
dr dr
To calculate the heat flow from the inner surface to the outside surface, we reorganize
Eq. (5.9) and then integrate thus,
dr
qr = −k.2πLdT
r
Z r2 Z T2
dr
qr = −k.2πL dT
r1 r T1
qr (ln r2 − ln r1 ) = −2πLk(T2 − T1 )
r2
qr ln = 2πLk(T1 − T2 )
r1
2πLk(T1 − T2 )
qr = r2 (5.10)
ln
r1

46
Note that unlike for a plane wall where the temperature distribution is linear, the
temperature distribution associated with radial conduction through a cylinder wall is
logarithmic.

5.1.3 Composite cylinder

Figure 5.4: Composite cylinder

Consider the composite wall of Fig. 5.4 and let the length of the cylinder = L.
Thermal resistance is calculated from Eq. (5.10) as
r2
ln
T1 − T2 r1
Rconduction = = (5.11)
qr 2πLk

47
From Fig. 5.4,
ln rr21 ln rr32 ln rr43
RA = , RB = , RC =
2πLkA 2πLkB 2πLkC
Therefore,
T1 − T4 T1 − T4
qr = = (5.12)
ΣR RA + RB + RC

5.1.4 The heat diffusion equation


A major objective in conduction analysis is to determine the temperature distribution,
representing how temperature varies with position in a medium. Once the temperature
distribution is known,

1. the conduction heat flux at any point in the medium or on its surface, can be
computed from Fourier’s law.

2. the structural integrity of a solid can be ascertained through determination of


thermal stresses, expansions and deflections.

3. the thickness of insulating material can be optimized.

4. compatibility of special coatings or adhesives used with the material can be de-
termined.

5.1.4.1 Determination of the temperature distribution


CARTESIAN COORDINATES:

1. Consider a differential control volume measuring dx by dy by dz, of a homoge-


neous material within which there is no advection (bulk motion); and in which
temperature distribution is experienced in cartesian co-ordinates T(x,y,z), Fig.
5.5

2. Consider the energy processes that are relevant to this control volume:

• if there are temperature gradients, conduction heat transfer occurs across


the control surfaces
• if the conduction heat rates perpendicular to the control surfaces at x, y and
z co-ordinate locations are qx , qy and qz respectively, the conduction heat
rates perpendicular to the opposite surfaces are expressed as a Taylor series

48
Figure 5.5: Control volume for heat diffusion equation in Cartesian coordinates

expansion (neglecting higher order terms) as:


∂qx
qx+dx = qx + dx (5.13)
∂x
∂qy
qy+dy = qy + dy (5.14)
∂y
∂qz
qz+dz = qz + dz (5.15)
∂z
• if there is an energy source within the medium, the rate of thermal energy
generated is expressed as,

Ėg = q̇dxdydz (5.16)

where q̇ is the rate at which energy is generated per unit volume of the
medium (units: W/m3 ).
NOTE: This term is positive (source) if thermal energy is generated in the
material and negative (sink) if thermal energy is consumed in the material.
• if energy is stored in the material, the rate at which this energy is stored is
expressed,
∂T
E˙st = ρcp dxdydz (5.17)
∂t
∂T
where ρcp is the time rate of change of the sensible thermal energy of
∂t
the medium per unit volume.

49
3. Apply the conservation of energy requirement

E˙in + Ėg − Eout


˙ = E˙st (5.18)
(qx + qy + qz ) + (q̇dxdydz) − (qx+dx + qy+dy + qz+dz )
∂T
= ρcp dxdydz (5.19)
∂t
NOTE:

(a) The energy generation and energy storage terms are volumetric phenomena
(b) The inflow and outflow terms are surface phenomena; they are associated
with processes occurring at the control surfaces and are proportional to the
surface areas

Substituting from Eqs. (5.13)-(5.15),

∂qx ∂qy ∂qz ∂T


− dx − dy − dz + q̇dxdydz = ρcp dxdydz (5.20)
∂x ∂y ∂z ∂t

From Fourier’s law; and with respect to Fig. 5.5,


∂T ∂T ∂T
qx = −kdydz ; qy = −kdxdz ; qz = −kdxdy (5.21)
∂x ∂y ∂z

substituting into Eq. (5.20), and dividing through by dxdydz,

∂ ∂T ∂ ∂T ∂ ∂T ∂T
(k )+ (k ) + (k ) + q̇ = ρcp (5.22)
∂x ∂x ∂y ∂y ∂z ∂z ∂t

This is the general form (in Cartesian coordinates) of the heat diffusion equation,
otherwise called the heat equation. It states that, at any point in the medium, the
rate of energy transfer by conduction into the unit volume plus the volumetric rate
of thermal energy generation, must equal the rate of change of thermal energy stored
within the volume.
Through solution of Eq. (5.22), the temperature distribution T (x, y, z) can be
obtained as a function of time.
5.1.4.2 Simplifications
1. If the thermal conductivity, k is constant (homogeneous material), we obtain,

∂ 2T ∂ 2T ∂ 2T q̇ 1 ∂T
2
+ 2
+ 2 + = (5.23)
∂x ∂y ∂z k α ∂t
k
α= is the thermal diffusivity - a physical property of the material
ρcp

50
2. Under steady state conditions (no change with respect to time), the right-hand
side of Eq. (5.22) reduces to zero

∂ ∂T ∂ ∂T ∂ ∂T
(k )+ (k ) + (k ) + q̇ = 0 (5.24)
∂x ∂x ∂y ∂y ∂z ∂z

3. If the heat transfer is one-dimensional, say in the x-direction only, Eq. (5.24)
simplifies to,
∂ ∂T
(k ) + q̇ = 0 (5.25)
∂x ∂x
and in the absence of heat generation, Eq. (5.25) further reduces to,

∂ ∂T
(k )=0 (5.26)
∂x ∂x

4. For a homogeneous material (k is constant), steady state conditions and without


heat generation, Eq. (5.22) simplifies to,

∂ 2T ∂ 2T ∂ 2T
+ + = 0, or ∇2 T = 0 (5.27)
∂x2 ∂y 2 ∂z 2
This is the Laplace equation.

CYLINDRICAL COORDINATES:
Given,

Figure 5.6: Control volume for heat diffusion equation in Cylindrical coordinates

51
Ėin = qr + qφ + qz
Ėout = qr+dr + qφ+dφ + qz+dz
Ėg = q̇(drdzrdφ)
∂T
Ėst = ρcp drdzrdφ
∂t
∂T
qr = −kdzrdφ
∂r
∂T
qφ = −kdrdz
rdφ
∂T
qz = −kdrrdφ
∂z
∂qr
qr+dr = qr + dr
∂r
∂qφ
qφ+dφ = qφ + rdφ
rdφ
∂qz
qz+dz = qz + dz
∂z
Then, applying the conservation of energy requirement, Eq. (5.18), it can be shown
that (students to do),
k ∂T ∂ ∂T ∂ ∂T ∂ ∂T ∂T
+ (k )+ (k ) + (k ) + q̇ = ρcp (5.28)
r ∂r ∂r ∂r r∂φ r∂φ ∂z ∂z ∂t
and for a homogeneous material (constant k ),

∂ 2T 1 ∂ 2T ∂ 2T
 
1 ∂T ∂T
k + 2 + 2 2 + 2 + q̇ = ρcp (5.29)
r ∂r ∂r r ∂φ ∂z ∂t

which for a steady state, one dimensional case (varying only in the radial direction),
and without heat generation (q̇ = 0),

1 dT d2 T
+ 2 = 0 (5.30)
r dr dr
For the cylinder of Fig. 5.3, the boundary equations are;
At r = r1 , T = T1
At r = r2 , T = T2
Thus, Eq. (5.30) can be integrated twice to give the temperature distribution.

5.2 Convection
Convection heat transfer occurs in fluids. It takes place by two mechanisms; random
molecular motion (diffusion) and bulk motion of the fluid. In convection, a portion
of the fluid mixes with another due to the gross motion of the mass of the fluid. If
the fluid motion is caused by density differences which are created by the temperature

52
difference existing in the mass of fluid, the process is termed free or natural convection
e.g., circulation of the water in a pan heated on a stove.
If the fluid motion is caused by external mechanical means e.g. by a fan or pump,
the convection is termed forced convection. An example of forced convection is use of
a fan to provide forced convection air cooling of hot electrical components on a printed
circuit board.
Regardless of the nature of the convection heat transfer process, the rate equation
is:
q 0 = h(Ts − T∞ ) (5.31)

where,
q 0 = convective heat flux (W/m2 )
Ts = surface temperature
T∞ = fluid temperature
h = convective heat transfer coefficient (W/m2 K).
The convective rate equation is also termed the Newton’s Law of Cooling. We will
analyse convection heat transfer as a boundary condition for conduction problems.
The heat rate by convection is determined from Eq. (5.31) as

q = q 0 A or q = hA(Ts − T∞ ) (5.32)

The thermal resistance for convection is then,


Ts − T∞ 1
Rconvection = = (5.33)
q hA

5.2.1 Overall heat transfer coefficient


We shall revisit the composite wall and besides conduction, we include convection.
Consider the one-dimensional heat transfer across the composite wall (Fig. 5.7),
bounded by a hot fluid on one end and a cold fluid on the other end.
The heat rate can be expressed,
T∞,1 − T∞,4
qx = (5.34)
ΣR
T∞,1 −T∞,4 is the overall temperature difference while ΣR is the total thermal resistance,
given by,
1 LA LB LC 1
ΣR = + + + + (5.35)
h1 A kA A kB A kC A h4 A
Since heat rate, qx is constant, then,
T∞,1 − Ts,1 Ts,1 − T2 T2 − T3
qx = = = =... (5.36)
R1 RA RB

53
Figure 5.7: Composite wall - overall heat transfer coefficient

It is often convenient to work with an overall heat transfer coefficient, U such that,
4T
qx = U A4T = , (5.37)
ΣR
therefore,
1 1
U = = (5.38)
AΣR 1 LA LB LC 1
+ + + +
h1 kA kB kC h4
In general,
4T 1
ΣR = = (5.39)
q UA
For the case of a multilayer cylinder, Fig. 5.8,
T∞,1 − T∞,4
qr = (5.40)
ΣR
T∞,1 − T∞,4
= r2 r3 r4 (5.41)
ln( ) ln( ) ln( )
1 r1 r2 r3 1
+ + + +
2πr1 Lh1 2πkA L 2πkB L 2πkC L 2πr4 Lh4

54
Figure 5.8: Composite cylinder with convection

In terms of the overall heat transfer coefficient,


T∞,1 − T∞,4
qr = = U A(T∞,1 − T∞,4 ) (5.42)
ΣR
1
∴ = UA (5.43)
ΣR
1
= U1 A1 = U2 A2 = U3 A3 = U4 A4 (5.44)
ΣR

55
1
For instance, from U1 A1 = , we can determine U1 thus,
ΣR
 
1 1 1
U1 = = , (5.45)
A1 ΣR 2πr1 L ΣR
i.e., the overall heat transfer coefficient based on the inside radius of the composite
cylinder. For this case, the heat rate can be expressed,

qr = U1 A1 (T∞,1 − T∞,4 ). (5.46)

5.3 Critical thickness of insulation


The above relations are applied in the insulation of pipes or electrical wires. As insu-
lation is added, the outer exposed surface temperature will decrease, but at the same
time, the surface area available for convective heat dissipation will increase. These two
opposing effects produce some interesting optimization effects. Referring to Fig. 5.9,
let the pipe radius be R and the insulation radius be r, so that (r-R) represents the
insulation thickness. For fixed values of the temperature of the fluid carried by the
pipe and the ambient air at Ta , the addition of insulation will alter the pipe surface
temperature, T (though the variation is so slight it may as well be neglected). Let the
thermal conductivity of the insulating material be k and the heat transfer coefficient
at the outside exposed insulation surface be h. Heat transfer from the pipe per unit
length is,
4T (T − Ta )
q = = r (5.47)
ΣR ln
1 R
+
2πLhr 2πLk
At the critical radius of insulation rc , the heat loss is a maximum while the thermal
resistance ΣR is a minimum. Increasing r beyond rc increases the thermal resistance
of the insulating layer. To obtain an expression for rc , we differentiate the thermal
resistance (i.e., the denominator of Eq. (5.47)) with respect to r, and equate to zero
(Students to do this. Hint: 2πL, h, k and R are all constant). Thus,
k
rc = (5.48)
h
HOMEWORK 2
A bare 2.5 cm-diameter pipe has a surface temperature of 1750 C and is placed in air
at 300 C. The convective heat transfer coefficient between the surface and the air is 5.6
W/m2 0 C. It is desired to reduce the heat loss to 50% of its present value by the addition
of an insulation with k =0.17 W/m0 C. Assuming that the pipe surface temperature and

56
Figure 5.9: Pipe insulation

the exposed surface heat transfer coefficient remain unchanged as insulation is added,
find the required thickness of insulation. Is this thickness an economically reasonable
value?

5.4 Radiation
Radiation is a term applied to many processes which involve energy transfer by elec-
tromagnetic wave phenomenon. We will limit our analysis to thermal radiation, i.e.
radiation produced by or which produces thermal excitation of a body. Thermal radi-
ation lies between 0.1 to 100µm of the electromagnetic spectrum. Radiation incident
on a surface may be reflected, absorbed or transmitted.

5.4.1 Definitions
1. Spectral dependency: Radiation incident on a surface may depend on the wave-
length. This is termed spectral dependency. The term monochromatic is used
to qualify a radiative quantity as applicable to only a single wavelength e.g.
monochromatic emission. The term total indicates evaluation over the entire
thermal radiation spectrum e.g. total emission.

2. Emissive power : This denotes the emitted thermal radiation leaving a surface, per
unit time, per unit surface area. It consists of only the original emission and does
not include any energy leaving a surface as a result of reflections. Monochromatic

57
emissive power, Eλ , is defined as the rate per unit area, at which a surface emits
thermal radiation at a particular wavelength, λ . The total emissive power, E,
and monochromatic emissive powers are related thus,
Z ∞
E = Eλ dλ (5.49)
0

3. Radiosity: This denotes all the radiation leaving a surface, per unit time, per
unit surface area. It differs from the emissive power in that it includes reflected
energy as well as the original emission. Similarly, the total radiosity, J, and the
monochromatic radiosity, Jλ , are related thus,
Z ∞
J = Jλ dλ (5.50)
0

4. Irradiation: This is the term used to denote the rate, per unit area, at which
thermal radiation is incident upon a surface (from all directions). It is a re-
sult of emissions and reflections form other surfaces, and may thus be spectrally
dependent. The total and monochromatic irradiation are related,
Z ∞
G = Gλ dλ (5.51)
0

5. Absorptivity: This is the fraction of the radiation incident on a surface, that is


absorbed by the surface.
αλ = monochromatic absorptivity: fraction of radiation absorbed at wavelength
λ.
α = total absorptivity: fraction absorbed at all wavelengths

6. Reflectivity: This is the fraction of the radiation incident on a surface, that is


reflected away from the surface.
ρλ = monochromatic reflectivity: fraction reflected at wavelength λ.
ρ = total reflectivity: fraction reflected at all wavelengths

7. Transmissivity: This is the fraction of the radiation incident on a surface that is


transmitted through the surface.
τλ = monochromatic transmissivity: fraction transmitted at wavelength λ
τ = total transmissivity: fraction transmitted at all wavelengths.
NOTE:
ρλ + αλ + τλ = 1, and ρ + α + τ = 1.

58
Figure 5.10: Reflectivity, absorptivity and transmissivity of incident radiation

These properties are dependent on the surface composition, roughness and tem-
perature. Air at atmospheric pressure is virtually transparent to thermal radiation so
that α ≈ ρ ≈ 0 and τ ≈ 1. Most solids (except glass), encountered in engineering are
opaque to thermal radiation such that τ = 0.
For thermally opaque solid surfaces,

ρ + α = 1,

and from the definitions of emissive power, radiosity and irradiation,

J = E + ρG
= E + (1 − α)G (5.52)

5.4.2 Blackbody and Graybody radiation


A perfect black body is an ideal body that absorbs all incident radiation regardless
of the spectral distribution or directional character of the incident radiation, that is,
αλ = α = 1, ρλ = ρ = 0.
Since a black body absorbs all incident radiation, the only radiation leaving a black
body surface is original emission.
The emissive power, Eb , of a black body depends on the surface temperature only
and is expressed by the Stefan-Boltzmann Law as,

Eb = σTs4 , (5.53)

59
where σ is the Stefan-Boltzmann constant with the value, σ = 5.67 x 10−8 W/m2 0 K4 .
Ts is the absolute temperature of the surface. The ratio of the emissive power of a
surface to that of a black body at the same temperature is called emissivity, ε,
E
ε = , (5.54)
Eb
that is, the emissivity of a surface is the measure of how closely a surface
approximates a blackbody, for which ε = 1. The value of emissivity for any
surface is in the range 0 ≤ ε ≤ 1. Note that when the emissivity of a non-black surface
is constant at all temperatures and through the entire range of wavelengths, the surface
is referred to as a gray body.
The maximum rate of radiation that can be emitted from a surface at a thermody-
namic temperature Ts is given as

Q̇ = σAs Ts4 [Watts]. (5.55)

This maximum rate of radiation applies to a black body. For any other surface (non-
black), the rate of radiation emitted from it is a function of its emissivity and is
expressed,

Q̇ = εσAs Ts4 [Watts]. (5.56)

When a surface of emissivity ε and surface area As at a thermodynamic temperature


Ts is completely enclosed by a much larger surface at a thermodynamic temperature
Tsurr , separated by a gas (such as air) that does not interfere with radiation, the net
rate of radiative heat transfer between these two surfaces is given by

Q̇rad = εσAs (Ts4 − Tsurr


4
) [Watts]. (5.57)

The law relating emissivity and absorptivity is called Kirchoff ’s law of radiation
and states that the emissivity and absorptivity of a surface at a given temperature and
wavelength are equal.

5.4.3 Spectral distribution of blackbody radiation


The total emissive power of an ideal black body is given by the Stefan-Boltzmann Law,
but it is also important to know the distribution of the emission from a black body over
the complete spectrum of wavelengths. This is given by the monochromatic emissive
power,

Ebλ = Ebλ (λ, T ), (5.58)

60
such that,
Z ∞
Eb = Ebλ (λ, T )dλ = σTs4 (5.59)
0

The dependence of Ebλ on λ and T is given by Planck’s Law as,


C1
Ebλ (λ, T ) = C2 (5.60)
λ5 (e λT − 1)
where C1 =3.7413 x 108 W-µm4 /m2 , and C2 =1.4388 x 104 µm4 K. Figure 5.11 shows a

Figure 5.11: Planck distribution for various temperatures

plot of Planck’s equation for several different temperatures. To determine the wave-
length at which Ebλ is a maximum, Planck’s equation is differentiated with respect to
λ and equated to zero. This leads to Wien’s Displacement Law,

λmax T = C3 , (5.61)

where the value of the constant C3 is 0.0028978 m 0 K.


This law may be stated thus, ”the product of absolute temperature and the

61
wavelength at which the emissive power is maximum, is constant”
EXAMPLES 9

1. A black body has a surface temperature of 5000 C. Determine the wavelength of


the maximum monochromatic emissive power

2. Consider a person standing in a breezy room at 200 C. Determine the total rate
of heat transfer from this person if the exposed surface area and the average
outer surface temperature of the person are 1.6 m2 and 290 C respectively, and
the convection heat transfer coefficient is 6 W/m2 K. Take the emissivity of the
person to be ε = 0.95 [168 W]

Tutorial Problems 7

1. A furnace wall consists of 125 mm wide refractory brick and 125 mm wide in-
sulating firebrick separated by an air gap. The outside wall is covered with a
12 mm thickness of plaster. The inner surface of the wall is at 11000 C and the
room temperature is 250 C. Calculate the rate at which heat is lost per m2 of
wall surface. The heat transfer coefficient from the outside wall surface to the
air in the room is 17 W/m2 K, and the resistance to heat flow of the air gap is
0.16 K/W. The thermal conductivity of refractory brick, insulating firebrick, and
plaster are 1.6, 0.3, and 0.14 W/mK, respectively. Calculate also each interface
temperature, and the temperature of the outside surface of the wall. [1.344 kW;
9950 C;7800 C;2200 C;1040 C]

2. A steel pipe of 100 mm inside diameter and 7 mm wall thickness, carrying steam
at 2600 C, is insulated with 40 mm of a moulded high-temperature diatomaceous
earth covering. This covering is in turn insulated with 60 mm of asbestos felt.
If the atmospheric temperature is 150 C, calculate the rate at which heat is lost
by the steam per m length of pipe. The heat transfer coefficients for the inside
and outside surfaces are 550 and 15 W/m2 K, respectively, and the thermal con-
ductivities of steel, diatomaceous earth and asbestos felt are 50, 0.09 and 0.07
W/mK respectively. Calculate also the temperature of the outside surface. [116
W; 22.80 C]

3. A 3-mm-diameter and 5-m-long electric wire is tightly wrapped with a 2-mm-


thick plastic cover whose thermal conductivity is k = 0.15 W/m.0 C. Electrical
measurements indicate that a current of 10 A passes through the wire and there is
a voltage drop of 8 V along the wire. If the insulated wire is exposed to a medium
at T∞ = 300 C with a heat transfer coefficient of h = 12 W/m2 .0 C, determine

62
(a) the total thermal resistance for the configuration [0.9377 0 C/W]
(b) the temperature at the interface of the wire and the plastic cover [105 0 C]

4. If a black body at 1000 K and a gray body at 1250 K emit the same amount of
radiation, what should be the emissivity of the gray body? [0.4096]

5. The filament of a 75 W light bulb may be considered a blackbody radiating into


a black enclosure at 700 C. The filament diameter is 0.1 mm and the length is 5
cm. Considering only radiation, determine the filament temperature [3029 K]

5.5 Heat Exchangers


A heat exchanger is a device that allows heat exchange between flowing fluids normally
separated by a dividing wall. The temperature of each fluid changes as it passes through
the exchanger, and hence the temperature of the dividing wall between the fluids also
changes along the length of the exchanger. Condensers and boilers in a steam plant
operate on this principle.
An engineer in practice often finds himself faced with the task of selecting a suitable
heat exchanger for a particular application, normally either:

• To achieve a specified temperature change in a fluid stream of known mass flow


rate, or

• To predict the outlet temperatures of the hot and cold fluid streams in a specified
heat exchanger.

In the first case, the log mean temperature difference method is best suited, while
for the second case, the effectiveness-NTU method is preferred.

5.5.1 Types of Heat Exchangers


There are three main types:

1. Recuperator: The flowing fluids that exchange heat are on either side of a
dividing wall. This is the most commonly used type in engineering practice and
ordinarily the term heat exchanger invariably refers to this type.

2. Regenerator: The hot and cold fluids pass alternately through a space contain-
ing solid particles. These particles provide alternately a sink and a source for
heat flow.

3. Evaporative type: A liquid is cooled evaporatively and continuously in the


same space as the coolant, for example, the cooling tower.

63
5.5.2 Types of Recuperative Heat Exchangers
5.5.2.1 Parallel Flow and Counter Flow Recuperators
The parallel flow configuration consists of a straight pipe through which one fluid
flows, while the second fluid flows in the same direction through an annulus surrounding
the pipe. In the counter flow configuration, the two fluids flow in opposite directions.
Let the mean inlet and outlet temperatures of the hotter fluid be Th,in and Th,out
respectively, and the corresponding temperatures for the colder fluid at inlet and outlet
be Tc,in and Tc,out respectively as represented on Fig. 5.12(a) (parallel flow) and
5.12(b) (counter flow). Let the mass flow rates of the hotter and colder fluids be ṁh
and ṁc respectively, and let their corresponding specific heat capacities (at constant
pressure) be cph and cpc respectively.

(a) (b)

Figure 5.12: Variations of the fluid temperatures for (a) parallel flow (b) counter flow
heat exchangers

64
For the parallel flow configuration, the temperature difference at the inlet is

4T1 = Th,in − Tc,in , (5.62)

while the temperature difference at the outlet (section 2) is

4T2 = Th,out − Tc,out . (5.63)

An overall heat transfer coefficient, U , can be defined for the thin tube wall separating
the fluids, based on Eq. (5.43), that is
1
U= , (5.64)
RT As
where As is the mean surface area of the tube, and RT the total thermal resistance.
Since the resistance of the tube wall is negligibly small, we consider the thermal resis-
tance to be made up of only the convective thermal resistance either side of the tube
wall thus,
1 1 1
= + . (5.65)
U hh hc
The values of hh and hc may vary along the length of the tube in practice, but suitable
mean values are used. A mean value of the overall heat transfer coefficient U is also
assumed.
Let us consider a differential section along the length of the heat exchanger between
plane A and plane B (refer to Fig. 5.12(a)). The temperature difference between the
hot and the cold fluid along plane A is,

(4T )A = (Th )A − (Tc )A , (5.66)

while the corresponding temperature difference along plane B is,

(4T )B = (Th )B − (Tc )B . (5.67)

Further, subtracting Eq. (5.67) from Eq. (5.66) gives

(4T )A − (4T )B = d(4T )


= [(Th )A − (Th )B ] − [(Tc )A − (Tc )B ]
= dTh − dTc
∴ d(4T ) = dTh − dTc . (5.68)

Note that the hotter fluid decreases in temperature by dTh along the differential ele-
ment, while the colder fluid increases in temperature by dTc .

65
A small amount of heat dQ̇ is transferred across the differential section, which can
be evaluated using Eq. (5.37) as

dQ̇ = U dAs 4T. (5.69)

The heat given up by the hotter fluid must equal the heat received by the colder fluid

dQ̇ = −ṁh cph dTh = ṁc cpc dTc , (5.70)

therefore
−dQ̇ dQ̇
dTh = and dTc = . (5.71)
ṁh cph ṁc cpc

Substituting Eq. (5.71) in Eq. (5.68)


 
−dQ̇ dQ̇ 1 1
d(4T ) = − = −dQ̇ + , (5.72)
ṁh cph ṁc cpc ṁh cph ṁc cpc

and integrating between the inlet (section 1) and the outlet (section 2)
 
1 1
4T1 − 4T2 = Q̇ + . (5.73)
ṁh cph ṁc cpc

Turning again to Eq. (5.72),

−d(4T )
dQ̇ =  . (5.74)
1 1
+
ṁh cph ṁc cpc

Substituting in Eq. (5.69)

−d(4T )
  = U dAs 4T, (5.75)
1 1
+
ṁh cph ṁc cpc
therefore
 
−d(4T ) 1 1
= dAs U + . (5.76)
4T ṁh cph ṁc cpc
Integrating between sections 1 and 2
 
4T2 1 1
−ln = As U + . (5.77)
4T1 ṁh cph ṁc cpc

From Eq.(5.73),
 
1 1 4T1 − 4T2
+ = , (5.78)
ṁh cph ṁc cpc Q̇

66
and substituting in Eq. (5.77)

4T2 As U (4T1 − 4T2 )


−ln = , (5.79)
4T1 Q̇
therefore
As U (4T1 − 4T2 )
Q̇ = . (5.80)
4T1
ln
4T2

Comparing Eq. (5.80) and Eq. (5.37), we can then define a mean temperature differ-
ence, 4Tmean , such that
4T1 − 4T2
4Tmean = . (5.81)
4T1
ln
4T2
This mean temperature difference is known as the logarithmic mean temperature dif-
ference, frequently given the symbol LMTD, such that

Q̇ = U · As · LM T D, (5.82)

where the mean surface area of the tube, As = πDl (D being the mean diameter of
the tube and l its total length). The analysis above applies to the counter flow heat
exchanger too, leading to a similar equation for the LMTD. The following notes are
important though.
NOTES

1. For parallel flow configuration,

4T1 = Th,in − Tc,in


4T2 = Th,out − Tc,out . (5.83)

For counter flow configuration,

4T1 = Th,in − Tc,out


4T2 = Th,out − Tc,in . (5.84)

2. In counter flow configuration, the temperature range is greater since theoretically,


the fluid being heated (colder fluid) can be raised to a higher temperature than
that of the heating fluid at the exit. In parallel flow configuration, the final
temperatures of the fluids must be somewhere between the initial values of each
fluid.

67
3. When the product ṁh cph is equal to ṁc cpc , then the temperature difference in
counter flow is the same all along the length of the tube, that is,

ṁh cph (Th,in − Th,out ) = ṁc cpc (Tc,in − Tc,out ),


(Th,in − Th,out ) = (Tc,in − Tc,out ),
(Th,in − Tc,in ) = (Th,out − Tc,out ),
4T = 4T1 = 4T2 .

This would however invalidate the use of Eq. (5.80), but application of l’Hôpital’s
rule reveals that the LM T D = 4T1 = 4T2 in that case.

4. For specified inlet and outlet temperatures, the LMTD is always greater for
counter flow than it is for parallel flow. Therefore, a smaller surface area (and
thus a smaller heat exchanger) is needed to achieve a specified heat transfer rate
in a counter flow heat exchanger. It is common practice then to use counter flow
arrangements in heat exchangers.
EXAMPLE 10
Exhaust gases flowing through a tubular heat exchanger at the rate of 0.3 kg/s are
cooled from 4000 C to 1200 C by water initially at 100 C. The specific heat capacities
of exhaust gases and water may be taken as 1.13 and 4.19 kJ/kg.K respectively, and
the overall heat transfer coefficient from gases to water is 140 W/m2 K. Calculate the
surface area required when the cooling water flow is 0.4 kg/s,
(a) for parallel flow;
(b) for counter flow.
SOLUTION
The heat given up by the exhaust gases is equal to the heat taken up by the water
(subscripts g and w denote the properties of hot gases and water respectively),

Q̇ = ṁg cp,g (4T )g = ṁw cp,w (4T )w .

Therefore

0.3 x 1.13 x (400 − 120) = 0.4 x 4.19 x (T − 10) = 94.92 kW


94.92
T = + 10
0.4 x 4.19
= 66.60 C. (5.85)

For parallel flow,


4T1 − 4T2 (400 − 10) − (120 − 66.6)
LM T D = = = 169.29 K, (5.86)
4T1
 
400 − 10
ln ln
4T2 120 − 66.6

68
and the surface area is then,

Q̇ 94.92 x 103
As = =
U · LM T D 140 x 169.29
2
=4m (5.87)

For counter flow,


4T1 − 4T2 (400 − 66.6) − (120 − 10)
LM T D = = = 201.47 K, (5.88)
4T1
 
400 − 66.6
ln ln
4T2 120 − 10
and the surface area is,

Q̇ 94.92 x 103
As = =
U · LM T D 140 x 201.47
2
= 3.37 m . (5.89)

Tutorial Problem 8
A counter-flow double-pipe heat exchanger is to heat water from 200 C to 800 C at a
rate of 1.2 kg/s. The heating is to be accomplished by geothermal water available
at 1600 C at a mass flow rate of 2 kg/s. The inner tube is thin-walled and has a
diameter of 1.5 cm. If the overall heat transfer coefficient of the heat exchanger is 640
W/m2 .0 C, determine the length of the heat exchanger required to achieve the desired
heating. Take the specific heats of water and geothermal fluid to be 4.18 and 4.31
kJ/kg.0 C respectively. (109 m. This is too long to be practical. A better
alternative is to use a plate heat exchanger or a multipass shell-and-tube
heat exchanger with multiple passes of tube bundles)

5.5.3 Effectiveness-NTU Method


This method is normally applied in cases where the inlet temperatures and mass flow
rates of the fluids are known but the outlet temperatures are not. It is based on a
dimensionless parameter called the heat transfer effectiveness ε, which is defined

actual heat transfer rate Q̇


ε= = . (5.90)
maximum possible heat transfer rate Q̇max
The actual heat transfer rate is determined from an energy balance on the cold/hot
fluid and can be expressed

Q̇ = Cc (Tc,out − Tc,in ) = Ch (Th,in − Th,out ), (5.91)

where Cc = ṁc cpc and Ch = ṁh cph are the heat capacity rates of the cold and hot
fluids respectively.

69
In determining the maximum possible heat transfer rate in the heat exchanger, the
maximum temperature difference is first determined thus

4Tmax = Th,in − Tc,in . (5.92)

Further, the heat transfer rate in a heat exchanger will reach a maximum value when
either:

• The cold fluid is heated to the inlet temperature of the hot fluid, or

• The hot fluid is cooled to the inlet temperature of the cold fluid.

These two conditions are not normally attained simultaneously because usually, the
heat capacity rates of the two fluids are not identical, i.e., Cc 6= Ch . Therefore, the
fluid with the smaller heat capacity rate experiences a larger temperature change, and
is thus the first to experience the maximum temperature. Consequently, the maximum
possible heat transfer rate is expressed

Q̇max = Cmin 4Tmax


= Cmin (Th,in − Tc,in ), (5.93)

where Cmin is the smaller between Ch and Cc .


If the effectiveness of the heat exchanger is known, the actual heat transfer rate can
be determined without knowing the outlet temperatures of the fluids. The effectiveness
depends on the geometry and flow configuration, and therefore different types of
heat exchangers have different effectiveness relations. These relations typically involve
the dimensionless group U As /Cmin . This quantity is called the number of transfer
units (NTU), and is further expressed

U As U As
NT U = = , (5.94)
Cmin (ṁcp )min

Where U is the overall heat transfer coefficient and As is the heat transfer surface area
of the heat exchanger. Note that for specified values of U and Cmin , the value of NTU
(which is directly proportional to As ) is a measure of the heat transfer surface area
and so, the larger NTU is, the larger the heat exchanger.
In heat exchanger analysis, another dimensionless parameter called the capacity
ratio can be defined thus
Cmin
c= . (5.95)
Cmax

70
Figure 5.13: NTU relations for heat exchangers

Effectiveness relations have been developed for a large number of heat exchangers,
and the results exist either in tables or charts. An example of such a table appears in
Fig. 5.13.
Tutorial Problems 9

1. Repeat Tutorial Problem 8 using the effectiveness-NTU method (Use Fig. 5.13).

71
REFERENCES
[1] M. D. Burghardt, Engineering Thermodynamics. Harper Collins, 1993.

[2] Z. Warhaft, An Introduction to Thermal-Fluid Engineering (The Engine and the


Atmosphere). Cambridge University Press, 1997.

[3] T. Eastop and A. McConkey, Applied Thermodynamics for Engineering Technolo-


gists. Longman, 1993.

[4] M. Horsley, Engineering Thermodynamics. Chapman and Hall, 1993.

[5] I. Granet, Thermodynamics and Heat Power. Prentice-Hall, 1985.

[6] Cengel, Yunus A. and Boles, Michael A., Thermodynamics - An Engineering Ap-
proach. McGraw-Hill, 2008.

72

You might also like