You are on page 1of 70

EMG 2302: ENGINEERING

THERMODYNAMICS II
Lecture Notes: Third Year BSc Mechanical
Engineering

J. K. Tanui

May - August, 2013


TABLE OF CONTENTS

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

1.0 Vapour Power Cycles 3


1.1 The Carnot Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Absolute Temperature Scale . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 The Rankine Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Rankine Cycle With Superheat . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Isentropic Efficiencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5.1 Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5.2 Pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.6 Regenerative Rankine Cycle . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6.1 Open Feedwater Heaters . . . . . . . . . . . . . . . . . . . . . . 12
1.6.2 Closed Feedwater Heaters . . . . . . . . . . . . . . . . . . . . . 13
1.7 Binary Vapour Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.7.1 Mercury-Water binary cycle . . . . . . . . . . . . . . . . . . . . 15
1.7.2 Water-Ammonia (or other refrigerant) binary cycle . . . . . . . 15
1.8 Refrigeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.8.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.8.2 The Vapour-Compression Refrigeration Cycle . . . . . . . . . . 18
1.8.3 Coefficient of Performance . . . . . . . . . . . . . . . . . . . . . 19
1.8.4 The Reversed Carnot Cycle . . . . . . . . . . . . . . . . . . . . 19
1.8.5 The Ideal Vapour-Compression Refrigeration Cycle . . . . . . . 20
1.8.6 Refrigerants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.0 Availability 28
2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2 Exergy Associated with Kinetic and Potential Energy . . . . . . . . . . 29
2.3 Reversible Work and Irreversibility . . . . . . . . . . . . . . . . . . . . 30
2.4 Exergy of a Nonflow (Closed) System . . . . . . . . . . . . . . . . . . . 31
2.5 Exergy of a Flow Stream . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.6 Second Law Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.6.1 Conventional Efficiencies . . . . . . . . . . . . . . . . . . . . . . 35
2.6.2 Determination of Second Law Efficiency . . . . . . . . . . . . . 36

i
3.0 The Ideal Gas and Gaseous Mixtures 39
3.1 The Perfect gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.1 Specific heat capacities . . . . . . . . . . . . . . . . . . . . . . . 40
3.1.2 Joule’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1.3 Relationship between specific heat capacities . . . . . . . . . . . 41
3.1.4 Specific enthalpy of a perfect gas . . . . . . . . . . . . . . . . . 42
3.1.5 Ratio of specific heat capacities . . . . . . . . . . . . . . . . . . 42
3.2 Gaseous mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.1 Gibbs-Dalton’s law . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2.2 Amagat’s (or Leduc’s) law . . . . . . . . . . . . . . . . . . . . . 45
3.2.3 Molar mass and specific gas constant for a gas mixture . . . . . 46
3.2.4 Specific heat capacities for a gas mixture . . . . . . . . . . . . . 47

4.0 Psychrometry 49
4.1 Dry and Atmospheric Air . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2 Specific and Relative Humidity of Air . . . . . . . . . . . . . . . . . . . 50
4.3 Dew-point Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4 Adiabatic Saturation and Wet-bulb Temperatures . . . . . . . . . . . . 52
4.5 The Psychrometric Chart . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.6 Human Comfort and Air Conditioning . . . . . . . . . . . . . . . . . . 57
4.7 Air Conditioning Processes . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.7.1 Simple Heating and Cooling . . . . . . . . . . . . . . . . . . . . 58
4.7.2 Heating with Humidification . . . . . . . . . . . . . . . . . . . . 59
4.7.3 Cooling with Dehumidification . . . . . . . . . . . . . . . . . . . 59
4.7.4 Evaporative Cooling . . . . . . . . . . . . . . . . . . . . . . . . 61
4.7.5 Mixing of Airstreams . . . . . . . . . . . . . . . . . . . . . . . . 64

ii
LIST OF FIGURES

1.1 The Carnot engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4


1.2 Carnot T-s cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 A T-s diagram for a Carnot cycle using steam . . . . . . . . . . . . . 5
1.4 T-s diagram for the Rankine cycle . . . . . . . . . . . . . . . . . . . . 7
1.5 Rankine cycle with superheat . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Isentropic efficiencies . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.7 Rankine cycle with superheat . . . . . . . . . . . . . . . . . . . . . . 11
1.8 Ideal regenerative Rankine cycle with open FWH . . . . . . . . . . . 12
1.9 Ideal regenerative Rankine cycle with a closed FWH . . . . . . . . . . 14
1.10 Mercury/water binary cycle . . . . . . . . . . . . . . . . . . . . . . . 16
1.11 Binary geothermal power plant . . . . . . . . . . . . . . . . . . . . . . 16
1.12 Basic components of a refrigeration system . . . . . . . . . . . . . . . 18
1.13 Carnot refrigerator and T-s diagram of the reversed Carnot cycle . . . 20
1.14 Schematic and T-s diagram of the ideal vapour-compression refrigera-
tion cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.15 The p-h diagram of an ideal vapour-compression refrigeration cycle . 22
1.16 T-s diagram for the ideal vapour-compression refrigeration cycle . . . 23

3.1 Amagat’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4.1 Schematic and T-s diagram for the adiabatic saturation process of water 53
4.2 Schematic diagram of a psychrometric chart . . . . . . . . . . . . . . 56
4.3 (a) Simple heating (b) Heating and dehumidifying . . . . . . . . . . . 59
4.4 Cooling and dehumidification . . . . . . . . . . . . . . . . . . . . . . 60
4.5 Air-conditioning processes . . . . . . . . . . . . . . . . . . . . . . . . 61
4.6 Psychrometric chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.7 Adiabatic mixing of two airstreams . . . . . . . . . . . . . . . . . . . 65

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

1. Understand the operation of typical vapour power cycles and basic refrigeration
cycles.

2. Know the concept of availability and its relation to the quality of energy.

3. Familiarize with properties of gaseous mixtures and appreciate basic air condi-
tioning principles.

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

1. Analyze vapour power cycles and ideal vapour compression refrigeration cycle.

2. Apply concepts of availability to open and closed systems.

3. Analyze properties of non-reacting gaseous mixtures.

4. Analyze basic air conditioning processes using a psychrometric chart.

Course Description
Vapour power cycles: Rankine, improved Rankine cycles. Regenerative cycles. Bi-
nary cycles. Reversed Carnot cycle: Refrigeration effect, coefficient of performance.
Ideal vapour-compression refrigeration cycle.
Availability: Definition. Availability equation for closed systems. Availability equa-
tion for open systems. Introduction to availability computations.
Gaseous mixtures: Non-reactive mixtures; mole fraction analysis, mass fraction anal-
ysis, volume fraction analysis. Gibbs-Dalton law. Relations involving pressure, volume,
internal energy, enthalpy, entropy and specific heats of gaseous mixtures. Vapour pres-
sure and condensation. Avogadro’s law.
Psychrometry: Specific properties of moist air. Adiabatic saturation temperature.
Mixing air streams. Presentation of moist air processes on a psychrometric chart. Air
conditioning processes.

Prerequisite(s)
EMG 2206: ENGINEERING THERMODYNAMICS I

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

2. T. D. Eastop & A. McConkey (1993) Applied Thermodynamics for Engineering


Technologists, 5th Edition.

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

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


Press.

3. International Journal of Fluid & Thermal Engineering.

Important Requirements
1. A copy of the tables of Thermodynamic Properties (Steam Tables)

2. A Scientific Calculator

3. Psychrometry Chart (to be provided)

2
1.0 Vapour Power Cycles
According to the first law of thermodynamics, when a closed system undergoes a com-
plete 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
[1, 2].

1.1 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. 1.1, while the T-s diagram is
shown on Fig. 1.2.
The thermal efficiency of the cycle is,
Wnet Wnet
ηth = = (1.1)
Qin QH
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 )

Therefore,
QH − QC Qnet QC
ηth = = =1− , (1.2)
QH QH QH

3
Figure 1.1: The Carnot engine

Figure 1.2: Carnot T-s cycle

but
QC TC (s3 − s4 )
= but (s2 − s1 ) = (s3 − s4 ),
QH TH (s2 − s1 )
therefore,
TC
ηth = 1 − (1.3)
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 1
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%]

4
HOMEWORK 1
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 kW; 20.88 kW]
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. 1.3 [3].

Figure 1.3: A T-s diagram for a Carnot cycle using steam

1.2 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 (equation (1.2)) 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

5
by X such that

ηth = φ(XH , XC ) (1.4)

where φ is a function while XH and XC are the temperatures of the hot and cold
reservoirs respectively.
Combining equations (1.2) and (1.4),
Qnet
= φ(XH , XC ) (1.5)
QH
The function φ cannot be determined analytically, for it is entirely arbitrary and many
temperature functions can satisfy it [4]. Kelvin proposed that the temperature function
may be chosen thus,
Qnet XC
1− = (1.6)
QH XH
From equation (1.3), we have
TC
ηth = 1 −
TH
or,
Qnet TC
1− = (1.7)
QH TH
Comparing equation (1.6) and (1.7), 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 [2, 4]

1.3 The Rankine Cycle


The 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. 1.3,

net work = Wout − Win


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

6
therefore,

(h2 − h3 ) − (h1 − h4 ) h1 − h4
work ratio = = 1−
(h2 − h3 ) h2 − h3
compression work
= 1− (1.8)
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. 1.3. It is more convenient to allow the condensation process to
proceed to completion, i.e. to the saturated liquid line, as in Fig. 1.4.

Figure 1.4: T-s diagram for the Rankine cycle

The working fluid is now water at state 4, Fig. 1.4, 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. 1.3. When compression work is reduced,
then from equation (1.8) the work ratio is increased. This modified cycle represented
by Fig. 1.4 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. 1.3. From equation (1.1), the thermal efficiency of
the cycle is thus reduced.
For the Rankine cycle,

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

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

7
• Heat rejection from the steam in the condenser, Qout = h3 − h4

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


p2 and p3 = p4

1.4 Rankine Cycle With Superheat


Between states 2 and 3 of Fig. 1.4, 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. 1.5. A typical value of the temperature of the steam at state 2 is 500-

Figure 1.5: Rankine cycle with superheat

6000 C. Metallurgical limitations prevent attainment of higher values. A wide range of


pressures can be used though.
EXAMPLES 2

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

8
output from the turbine, neglecting the feed pump work [986 kJ/kg].
Without neglecting feed pump work, determine
(a) the heat supplied [2685 kJ/kg], and
(b) the Rankine efficiency [36.6%]

Tutorial Problems 1

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%]

1.5 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.

1.5.1 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,

9
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. 1.6,
h2 − h30
ηT = (1.9)
h2 − h3
Process 2 − 30 is the actual steam expansion in the turbine while process 2 − 3 is the
isentropic expansion.

Figure 1.6: Isentropic efficiencies

1.5.2 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. 1.6,
h1 − h4
ηC = (1.10)
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 [2].
Tutorial Problems 2
The maximum steam temperature is 5600 C and the lowest cycle temperature is 300 C.

10
1. What are the thermal efficiencies of ideal Rankine and Carnot cycles operating
with a maximum pressure of

(a) 3.5 MPa?


(b) 7.0 MPa?

2. For the same temperatures and pressures in (1) above, recalculate the thermal
efficiency for a Rankine cycle with a turbine efficiency of 80%

1.6 Regenerative Rankine Cycle


Let us re-examine the Rankine cycle with superheat (Fig. 1.7. Heat is transferred to
the working fluid during process 2-20 at a relatively low temperature thus lowering the
average heat addition temperature and cycle efficiency. This shortcoming is remedied

Figure 1.7: Rankine cycle with superheat

by raising the temperature of the liquid leaving the pump (feedwater) before it enters
the boiler. A practical way of achieving this is by extracting (0 bleeding0 ) steam from the
turbine at various points, for heating the feedwater. The device where the feedwater

11
is heated is called the regenerator or feedwater heater (FWH).
Regeneration:

• Improves cycle efficiency

• Provides a convenient way of deaerating the feedwater (removing the air that
normally leaks in at the condenser) thus preventing corrosion in the boiler

• Helps to control the large volume flow rate of the steam at the final stages of the
turbine (due to the large specific volumes at low pressures)

Essentially, a FWH is a heat exchanger where the heat is transferred from the steam
to the feedwater either by mixing the two fluid streams (open feedwater heaters) or
without mixing them (closed feedwater heaters).

1.6.1 Open Feedwater Heaters


These are direct-contact FWH and they are basically mixing chambers where the steam
extracted from the turbine mixes with the feedwater exiting the pump. An ideal
regenerative Rankine cycle with open FWH is shown in Fig. 1.8. Processes:

Figure 1.8: Ideal regenerative Rankine cycle with open FWH

• 5 - 6: Steam enters the turbine at the boiler pressure (state 5) then expands
isentropically to an intermediate pressure (state 6).

• 6 - 3: Some steam is extracted at state 6 and routed to the FWH. The extracted
steam mixes with the feedwater and what comes out is a saturated liquid at state
3.

12
• 6 - 7: The steam remaining in the turbine continues to expand isentropically to
the condenser pressure (state 7).

• 7 - 1: The steam at state 7 loses heat in the condenser and leaves as a saturated
liquid at state 1.

• 1 - 2: The condensed water (feedwater) at state 1 then enters an isentropic


pump, where it is compressed to the FWH pressure (state 2).

• 2 - 3: The feedwater is routed to the FWH and mixes with the steam extracted
from the turbine - the mixture comes out as a saturated liquid at state 3.

• 3 - 4: A second pump raises the pressure of the water at state 3 to the boiler
pressure (state 4).

• 4 - 5: The cycle is completed by heating the water in the boiler to the turbine
inlet state (state 5).

Pertinent equations are given next. Quantities are expressed per unit mass of steam,
such that for each 1 kg of steam leaving the boiler, y kg expands partially and is
extracted at state 6, while the remaining (1 - y) kg expands completely to the condenser
pressure.

qin = h5 − h4 , (1.11)
qout = (1 − y)(h7 − h1 ), (1.12)
wturb,out = (h5 − h6 ) + (1 − y)(h6 − h7 ), (1.13)
wpump,in = (1 − y)wpump I,in + wpump II,in , (1.14)
= (1 − y)[v1 (p2 − p1 )] + v3 (p4 − p3 ), (1.15)
ṁ6
y= (= fraction of steam extracted). (1.16)
ṁ5
The more the feedwater heaters used, the higher the cycle efficiency though the opti-
mum number is determined from economic considerations.

1.6.2 Closed Feedwater Heaters


Heat is transferred from the extracted steam to the feedwater without any mixing
taking place. In an ideal closed FWH, the feedwater is heated to the exit temperature
of the extracted steam, which ideally leaves the heater as a saturated liquid at the
extraction pressure. In actual power plants, the feedwater leaves the heater below the
exit temperature of the extracted steam because a temperature difference of at least a

13
Figure 1.9: Ideal regenerative Rankine cycle with a closed FWH

few degrees is required for any effective heat transfer to take place. Figure 1.9 shows
a schematic diagram of a steam power plant with one closed FWH.
While the open FWH is simple and inexpensive, a separate pump is required for
each open FWH used. The closed feedwater heaters are more complex in design and
are also more expensive, but do not require a separate pump for each FWH since the
extracted steam and the feedwater can be at different pressures.
ASSIGNMENT 1, Qn. 1 - Due on
Consider a steam power plant that operates on an ideal reheat-regenerative Rankine
cycle with one open feedwater heater, one closed feedwater heater, and one reheater.
Steam enters the high-pressure turbine at 15 MPa and 6000 C and expands to a pressure
of 4 MPa. At this pressure, some steam is extracted from the turbine for the closed
feedwater heater, and the remaining steam is reheated at the same pressure to 6000 C.
The extracted steam is completely condensed in the heater and is pumped to 15 MPa
before it mixes with the feedwater at the same pressure. Steam for the open feedwater
heater is extracted from the low-pressure turbine at a pressure of 0.5 MPa, while the
remaining steam is condensed in the condenser at a pressure of 10 kPa.

1. Represent this problem both on a schematic and also a T-s diagram

2. Neglecting KE and PE changes, and assuming that in both the open and closed
FWH the feedwater is heated to the saturation temperature at the FWH pressure,
determine

14
(a) The fractions of steam extracted from the turbines
(b) The thermal efficiency of the cycle

1.7 Binary Vapour Cycles


Although water is widely used as a working fluid in power plant cycles because of its
abundance, ease of handling and nontoxicity, it also has several disadvantages:

1. has a high vapour pressure, necessitating thick piping and thick pressure vessels,
thus presenting design difficulties

2. superheated steam is a poor heat transfer fluid thus requiring larger heat transfer
surfaces making equipment expensive

3. water vapour (steam) often changes phase from superheated to wet in the turbine,
leading to degradation of the turbine

4. has a high specific heat capacity

5. has to be conditioned in order to be used in the boiler, otherwise it is corrosive

For these reasons, water is also used in binary vapour cycles to utilize its known ad-
vantages in conjunction with another suitable fluid [4]. A binary cycle is actually a
combination of two cycles, one in the high-temperature region (this is the topping cy-
cle), and the other in the low-temperature region (this is called the bottoming cycle)
[5]. The heat output from the topping cycle is used as the heat input to the bottom-
ing cycle. Examples of working fluids suitable for the topping cycle include mercury,
sodium, potassium, and sodium-potassium mixtures [5].

1.7.1 Mercury-Water binary cycle


Mercury is used as the high temperature working fluid, and the heat rejected by the
mercury part of the cycle is used as the heat source for the steam part of the cycle
(Fig. 1.10).

1.7.2 Water-Ammonia (or other refrigerant) binary cycle


Such a plant uses hot water as the high temperature working fluid and a secondary
fluid of a lower boiling point than water, to produce the vapour for running the turbine
in a power plant (Fig. 1.11).
0
1. One application is in a binary geothermal power plant (geothermal means heat
0
from the earth ). This is applied in geothermal areas containing moderate-
temperature geothermal resource (water), i.e., below 4000 F.

15
Figure 1.10: Mercury/water binary cycle

The hot geothermal fluid and a secondary fluid of a much lower boiling point than
water pass through a heat exchanger. Heat from the geothermal fluid causes the
secondary fluid to flash to vapour, which then drives the turbine. Past secondary
working fluids were CFC (Freon type) refrigerants. Current plants use hydro-
carbons (isobutane, pentane etc) or HFC type refrigerants. The specific fluid is
chosen to match the geothermal resource temperature.

Figure 1.11: Binary geothermal power plant

2. Another application is the OTEC (Ocean Thermal Energy Conversion) plant.


This plant uses warm surface water of the ocean as the heat source. This warm
water vapourizes a liquid, such as ammonia, and the ammonia vapour drives

16
the turbine-generator. The ammonia vapour, after running the turbine, is then
condensed back to liquid by cold ocean water [4].

Tutorial Problems 3

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

2. (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.

3. 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]

4. 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]

1.8 Refrigeration
1.8.1 Definitions
Ordinarily, heat is transferred in the direction of decreasing temperature, that is, from
high-temperature mediums to low-temperature ones. This heat transfer process occurs
in nature without requiring the aid of any device [5]. The reverse process cannot occur
without the aid of a device. Thus, the transfer of heat from a low-temperature medium
to a high-temperature one requires special devices, one of which is a refrigerator. The
refrigerator maintains the refrigerated space at a low-temperature. Another device
that functions like a refrigerator is the heat pump, only that a heat pump discharges
heat from a low-temperature medium to a high-temperature one so as to maintain a
heated space at a high temperature, e.g., absorbing heat from cold outside air in
winter and supplying it to the warmer interior of a house.

17
Refrigerators are cyclic devices and the working fluids used in the refrigeration
cycles are called refrigerants. The most frequently used refrigeration cycle is the
vapour-compression refrigeration cycle.

1.8.2 The Vapour-Compression Refrigeration Cycle


This cycle is made up of four main components: A compressor, a condenser, an ex-
pansion valve and an evaporator, as represented in Fig. 1.12. Typical operating
conditions are also shown in the figure. The refrigerant enters the compressor as a

Figure 1.12: Basic components of a refrigeration system

vapour and is compressed to the condenser pressure. It leaves the compressor at a


relatively higher temperature then cools down and condenses as it flows through the
coils of the condenser, by rejecting heat to the surrounding medium. It then enters a
capillary tube where its pressure and temperature drop drastically due to throttling.
The low-temperature refrigerant then enters the evaporator where it evaporates by
absorbing heat from the refrigerated space. The cyclic process is completed by the
refrigerant leaving the evaporator and reentering the compressor.

18
1.8.3 Coefficient of Performance
The performance of a refrigerator is expressed in terms of the coefficient of performance,
COPR , defined thus
desired output cooling effect QL
COPR = = = , (1.17)
required input work input Wnet,in

where QL is the heat removed from the refrigerated space, and is also referred to as
the refrigeration effect. By the conservation of energy principle for a cyclic device,

Wnet,in = QH − QL , (1.18)

where QH is the heat rejected to the warm medium. Therefore,


QL 1
COPR = = . (1.19)
QH − QL QH /QL − 1

1.8.4 The Reversed Carnot Cycle


We recall from Section 1.1 that a Carnot cycle is a totally reversible cycle consisting of
two reversible isothermal and two isentropic processes. It has the maximum thermal
efficiency for given temperature limits and serves as a standard against which actual
power cycles can be compared.
If we reverse the Carnot cycle thereby also reversing the directions of the heat and
work interactions, the result is a reversed Carnot cycle. A refrigerator that operates on
the reversed Carnot cycle is called a Carnot refrigerator. Consider such a refrigerator,
as shown in Fig. 1.13 side by side with the T-s diagram of the reversed Carnot cycle.
The refrigeration cycle is then described:
Process 1-2: The refrigerant absorbs heat QL isothermally from a low-temperature
source at TL .
Process 2-3: The refrigerant is compressed isentropically to state 3, reaching a temper-
ature TH .
Process 3-4: The refrigerant rejects heat QH isothermally to a high-temperature sink
at TH . During this process, the refrigerant changes from a saturated vapour state to a
saturated liquid state.
Process 4-1: The refrigerant expands isentropically to state 1, dropping back to tem-
perature TL .
The coefficient of performance of a Carnot refrigerator is expressed in terms of the
temperatures as
1
COPR,Carnot = , (1.20)
TH /TL − 1

19
Figure 1.13: Carnot refrigerator and T-s diagram of the reversed Carnot cycle

and increases as the difference between the two temperatures decreases. The re-
versed Carnot cycle is the most efficient refrigeration cycle operating between two
specified temperatures, but is unsuitable in practice because of the following challenges:

1. Process 2-3 cannot be closely approximated in practice because it involves the


compression of a liquid-vapour mixture, which requires a compressor that will
handle two phases.

2. Process 4-1 also cannot be closely approximated in practice because it involves the
expansion of high-moisture-content refrigerant, which is detrimental to a turbine.

However, the other two processes, 1-2 and 3-4, can be approached closely in actual
evaporators and condensers respectively [5].

1.8.5 The Ideal Vapour-Compression Refrigeration Cycle


A great extent of the challenges associated with the reversed Carnot cycle can be
eliminated by:

1. Vapourizing the refrigerant completely before it is compressed.

2. Replacing the turbine with a throttling device, such as an expansion valve or


capillary tube.

20
The resulting cycle is the ideal vapour-compression refrigeration cycle, shown
schematically in Fig. 1.14.

Figure 1.14: Schematic and T-s diagram of the ideal vapour-compression refrigeration
cycle

The vapour-compression refrigeration cycle is the most widely used cycle not just
for refrigerators but also for air-conditioning systems and heat pumps. Other than the
T-s diagram, the p-h diagram is frequently used in the analysis of vapour-compression
refrigeration cycles, and is shown in Fig. 1.15. It is important to note that the
ideal vapour-compression refrigeration cycle is not an internally reversible cycle since
it involves an irreversible (throttling) process, which is maintained in the cycle to make
it a more realistic model for the actual vapour-compression refrigeration cycle.
The four components associated with the vapour-compression refrigeration cycle
(evaporator, compressor, condenser and expansion valve) are steady-flow devices mean-
ing that the corresponding processes can be analyzed as steady-flow. Since the KE and
PE changes of the refrigerant are small compared to the work and heat transfer terms,
they can be neglected such that the steady-flow energy equation on a unit-mass basis
simplifies to

(qin − qout ) + (win − wout ) = ho − hi (1.21)

21
Figure 1.15: The p-h diagram of an ideal vapour-compression refrigeration cycle

With reference to Fig. 1.15, we may express the coefficient of performance of a


refrigerator as
qL h1 − h4
COPR = = , (1.22)
wnet,in h2 − h1

where h1 = hg at p1 and h3 = hf at p3 for the ideal case.


EXAMPLE 3
A refrigerator uses refrigerant-134a as the working fluid and operates on an ideal
vapour-compression refrigeration cycle between 0.14 and 0.8 Mpa. If the mass flow
rate of the refrigerant is 0.05 kg/s, determine:
(a) The rate of heat removal from the refrigerated space and power input to the com-
pressor
(b) The rate of heat rejection to the environment, and
(c) The COP of the refrigerator
SOLUTION
The T-s diagram is shown in Fig. 1.16. From the refrigerant-134a tables, the en-

22
Figure 1.16: T-s diagram for the ideal vapour-compression refrigeration cycle

thalpies and entropies at the various states are:

h1 = hg at 0.14 MPa = 239.16 kJ/kg,


s1 = s2 = sg at 0.14 MPa = 0.94456 kJ/kg.K,
h2 = 275.39 kJ/kg,
h3 = hf at 0.8 MPa = 95.47 kJ/kg,
h4 ∼
= h3 (throttling) = 95.47 kJ/kg.

(a) The rate of heat removal:

Q̇L = ṁ(h1 − h4 ) = (0.05 kg/s)[(239.16 - 95.47)kJ/kg] = 7.18 kW

The power input to the compressor:

Ẇin = ṁ(h2 − h1 ) = (0.05 kg/s)[(275.39 - 239.16)kJ/kg] = 1.81 kW

Note that since no work transfer is associated with the expansion valve (i.e. h4 ∼
= h3 ),
then Ẇin = Ẇnet,in .

23
(b) The rate of heat rejection to the environment:

Q̇H = ṁ(h2 − h3 ) = (0.05 kg/s)[(275.39 - 95.47)kJ/kg] = 9.0 kW,

or alternatively

Q̇H = Q̇L + Ẇnet,in = 7.18 kW + 1.81 kW = 8.99 kW

(c) The coefficient of performance is

Q̇L 7.18 kW
COPR = = = 3.97
Ẇnet,in 1.81 kW

HOMEWORK 2

1. If the expansion valve in the problem above were to be replaced by an isen-


tropic turbine, what would be the effect on the coefficient of performance of the
refrigerator?

2. An actual vapour-compression refrigeration cycle differs from the ideal one in


several ways, owing to the irreversibilities that occur in the various components.
Discuss the actual vapour-compression refrigeration cycle.

ASSIGNMENT 1, Qn. 2 - Due on


Refrigerant-134a enters the compressor of a refrigerator as superheated vapour at 0.14
MPa and -100 C at a rate of 0.05 kg/s and leaves at 0.8 MPa and 500 C. The refriger-
ant is cooled in the condenser to 260 C and 0.72 MPa and is throttled to 0.15 MPa.
Disregarding any heat transfer and pressure drops in the connecting lines between the
components, determine
(a) The rate of heat removal from the refrigerated space and the power input to the
compressor
(b) The isentropic efficiency of the compressor, and
(c) The coefficient of performance of the refrigerator
Neglect kinetic and potential energy changes.

1.8.6 Refrigerants
1.8.6.1 Background
It is of paramount importance to select the right refrigerant for a particular applica-
tion. Some of the refrigerants that have been in use so far include chlorofluorocarbons
(CFCs), ammonia, hydrocarbons (propane, ethane, ethylene etc), carbon dioxide, air
(in the air-conditioning of aircraft) and even water (applications above freezing point).

24
Ethyl ether was the first commercially used refrigerant in vapour-compression
systems in 1850. This was then followed by ammonia, carbon dioxide, methyl chloride,
sulphur dioxide, butane, ethane, propane, isobutane, gasoline, and chlorofluorocarbons,
amongst others.
Ammonia, though toxic, has been a favourite over the years in a wide variety of
applications predominantly in large-scale food refrigeration facilities (cooling of fresh
fruits, vegetables, meat, fish, beer, wine etc) because of the following advantages:

1. Low cost

2. Higher comparative COP

3. More favourable thermodynamic and transport properties, thus higher heat trans-
fer coefficients

4. Greater detectability in case of leakages

5. No effect on the ozone layer

Its toxicity though makes it unsuitable for domestic use.


Early domestic refrigerants were also quite toxic and included sulphur dioxide, ethyl
chloride and methyl chloride. After some fatalities in the 1920s, it was imperative
to come up with alternative refrigerants for domestic use. That is when the CFCs
came into the market, notably R-21, then R-12 and R-11. The CFCs proved to be
versatile and of low cost and were also widely used in aerosols, foam applications, and
as solvents. R-12 was particularly outstanding in domestic applications. R-502, a blend
of R-115 and R-22, was a dominant refrigerant in commercial refrigeration systems e.g.
in supermarkets.
Then came the ozone crisis, when it was realized (in the mid-1970s) that CFCs al-
lowed more ultraviolet radiation into the earth’s atmosphere by destroying the protec-
tive ozone layer, thus contributing to the greenhouse effect that causes global warming.
The most harmful of the CFCs have been R-11, R-12 and R-115, with R-22 being less
harmful. Consequently, many international treaties have banned the use of CFCs. In
particular, the popular R-12 has largely been replaced by the chlorine-free R-134a.
1.8.6.2 Selection of Refrigerants
Two important parameters need to be considered in selecting the right refrigerant:

1. Temperature of the refrigerated space

2. Temperature of the environment

25
A temperature difference of 5 - 100 C needs to be maintained between the refrigerant
and the medium with which it exchanges heat. In addition, the refrigerant should
maintain a pressure of 1 atm or higher to avoid any leakage into the refrigeration
system.
Desirable characteristics of the refrigerant include:

• Non-toxicity

• Non-corrosiveness

• Non-flammability

• Chemical stability

• Low cost

• High enthalpy of vapourization (this minimizes mass flow rate)

In cases where one refrigerant does not meet all the desired requirements, two or more
refrigeration cycles operating with different refrigerants may be used in series, and the
resulting system is called a cascade system.
Tutorial Problems 4

1. A steady-flow Carnot refrigeration cycle uses refrigerant-134a as the working


fluid. The refrigerant changes from saturated vapour to saturated liquid at 300 C
in the condenser as it rejects heat. The evaporator pressure is 160 kPa. Show
the cycle on a T-s diagram relative to saturation lines, and determine
(a) The coefficient of performance [5.64]
(b) The amount of heat absorbed from the refrigerated space [147 kJ/kg]
(c) The net work input [26.1 kJ/kg]

2. An ideal vapour-compression refrigeration cycle that uses refrigerant-134a as its


working fluid maintains a condenser at 1000 kPa and the evaporator at 40 C.
Determine the system’s COP and the amount of power required to service a 400
kW cooling load [6.46, 61.9 kW]

3. A refrigerator uses refrigerant-134a as the working fluid and operates on the ideal
vapour-compression refrigeration cycle. The refrigerant enters the evaporator
at 120 kPa with a quality of 30% and leaves the compressor at 600 C. If the
compressor consumes 450 W of power, determine
(a) The mass flow rate of the refrigerant [0.00727 kg/s]

26
(b) The condenser pressure [672 kPa]
(c) The COP of the refrigerator [2.43]

4. An ammonia vapour-compression refrigerator operates between an evaporator


pressure of 2.077 bar and a condenser pressure of 12.37 bar. The following cycles
are to be compared by calculating the COP and the refrigeration effect per kg; in
each case, there is no undercooling in the condenser, and isentropic compression
may be assumed:
(a) The vapour has a dryness fraction of 0.9 at entry to the compressor [4.5;
957.5 kJ/kg]
(b) The vapour is dry saturated at entry to the compressor [4.13; 1089.9
kJ/kg] (c) The vapour has 5 K of superheat at entry to the compressor [4.1;
1101.4 kJ/kg]
What would be the COP of a reversed cycle operating between the same satura-
tion temperatures? [5.1]

27
2.0 Availability
2.1 Definition
We can combine the first and second laws of thermodynamics to determine theoretical
limits of performance of various thermodynamic components and systems. Thus, we
introduce the concept of availability, otherwise called exergy. This may simply be
defined as the maximum useful work that can be obtained from a system
at a given state in a specified environment.
The first law of thermodynamics involves the conservation of energy: No energy
can be created or destroyed in any energy conversion process. As such, all energy
can be accounted for during the conversion process. The first law is the basis of the
energy-balance method frequently used in the analysis and design of energy systems.
Thus, the first law deals with the QUANTITY of energy.
The second law of thermodynamics on the other hand deals with the QUALITY
of energy and has many ramifications with respect to energy conversion processes. It
enables the engineer to determine whether or not the desired conversion is possible.
Consider a coal-fired power plant in which coal energy is converted to electrical energy
with waste heat as a by-product. As per the second law of thermodynamics, we know
that the reverse process, that is, converting the electrical energy plus waste heat back
to coal energy, is impossible. While energy is conserved in the conversion process, its
quality, which is its ability to cause a change, deteriorates. In other words, while the
QUANTITY of energy is conserved in an irreversible energy conversion process (first
law is obeyed), the QUALITY of energy is degraded.
From the foregoing, availability is embodied in the second law and represents the
work-producing potential of the substance. And since not all energy has the same
ability to cause a change, one form of energy is not equivalent to another form even
though they may be numerically the same, for example: Steam from a boiler has
thermal energy and water behind a dam has potential energy. Though the two
forms of energy may be numerically equal, they will not necessarily have equal work-
producing potentials. In that case their availabilities would be different.
Though the wording may vary from author to author, availability (or exergy),
should be simply understood as referring to the maximum theoretical work that can
be extracted from a system in passing from the given state to a state of equilibrium
with the environment, i.e., maximum theoretical work extracted as the system changes
state to the dead state. Note that in the dead state, the system possesses energy but

28
no exergy. We shall denote properties of a system at the dead state by subscript zero.
Further, we shall assume (unless otherwise specified) that the dead-state temperature
and pressure are T0 = 250 C and p0 = 1 atm (or 101.325 kPa) respectively.

2.2 Exergy Associated with Kinetic and Potential


Energy
Kinetic and potential energies are forms of mechanical energy. These two forms can be
converted to work entirely [5].
The work potential or exergy of the kinetic energy of a system is equal to the
kinetic energy itself. Thus,
C2
xke = KE = , (2.1)
2
where C is the velocity of the system relative to the environment.
The exergy of the potential energy of a system is equal to the potential energy
itself. Thus,

xpe = PE = gz, (2.2)

where g is the gravitational acceleration and z is the elevation of the system relative
to a reference level in the environment.
The exergies of the kinetic and potential energies are equal to themselves and are
entirely available for work.
EXAMPLE 4
A wind turbine has a rotor of 12 m in diameter. It is to be installed at a location
where the wind is blowing steadily at an average velocity of 10 m/s. Determine the
maximum power that can be generated by the wind turbine, assuming the air to be at
standard conditions of 1 atm and 250 C where its density is 1.18 kg/m3 .
SOLUTION
We shall ignore Betz’s law and the accompanying dynamics of wind power generation.
We only invoke availability analysis in this case. The exergy (or availability) of the
blowing air is simply the kinetic energy it possesses, thus
C2 (10 m/s)2
KE (per unit mass) = = = 50 J/kg( or 0.05 kJ/kg). (2.3)
2 2
The mass flow rate, ṁ, of the air is
πD2 π(12 m)2
ṁ = ρAC = ρ C = (1.18 kg/m3 ) (10 m/s) = 1334.55 kg/s. (2.4)
4 4
The maximum power is the product of the mass flow rate and the KE,

ṁ x KE = (1334.55 kg/s)(0.05 kJ/kg) = 66.73 kW. (2.5)

29
2.3 Reversible Work and Irreversibility
Actual engineering systems hardly reach the dead state while in operation. Addition-
ally, the work done by a work-producing device is not always entirely in a usable form,
e.g., when a gas in a piston-cylinder expands, part of the work done by the gas is used
to push the atmospheric air out of the way of the piston. This work cannot be recovered
and utilized for a useful purpose. It represents an irreversibility,I, otherwise called
exergy-destruction. Irreversibility can be viewed as the wasted work potential and
represents the energy that could have been converted to useful work, Wu , but was not.
The performance of a system is enhanced by minimizing the irreversibility associated
with it.
On the other hand, reversible work, Wrev , is the maximum amount of useful work
that can be produced as a system undergoes a process between the initial and final
states. When the final state is the dead state, reversible work then equals exergy.
In a given process, any difference between the reversible work (Wrev ) and the useful
work (Wu ), is due to the irreversibilities present during the process. For a device that
gives out work, e.g. a heat engine, this may be expressed

I = Wrev,out − Wu,out , (2.6)

while for a device that consumes work, e.g. a compressor, we have

I = Wu,in − Wrev,in . (2.7)

Note that Wrev ≥ Wu for work-producing devices, and Wrev ≤ Wu for work-consuming
devices.
EXAMPLE 5
A heat engine receives heat from a source at 1200 K at a rate of 500 kW and rejects
the waste heat to a medium at 300 K. The power output of the heat engine is 180 kW.
Determine the reversible power and the irreversibility rate of this process.
SOLUTION
The reversible power is the amount that a reversible engine e.g., a Carnot engine, would
produce when operating between the same temperature limits, and is determined thus
   
Tsink 300 K
Ẇrev = 1 − Q̇in = 1 − (500 kW) = 375 kW. (2.8)
Tsource 1200 K
The irreversibility rate is the difference between the reversible power (maximum power
that could have been produced) and the useful power output

I˙ = Ẇrev,out − Ẇu,out = 375 − 180 = 195 kW. (2.9)

This means that 195 kW of the power potential is wasted during this process due to
irreversibilities.

30
2.4 Exergy of a Nonflow (Closed) System
We recall that the dead state of a system (whose properties are denoted by subscript
0) is a state in which the system is at equilibrium with the environment and therefore
possesses no exergy. If we limit ourselves to thermo-mechanical exergy (i.e., without
considering mixing and chemical reactions), the system in the dead state cannot do any
work and is at the pressure and temperature of the environment, and has no kinetic
or potential energy relative to the environment. Therefore, the total useful work that
a closed system would deliver in undergoing a reversible process from a given state to
the dead state would be the exergy of the closed system, X, which can be expressed in
terms of the internal, kinetic and potential energies thus,

C2
X = (U − U0 ) + p0 (V − V0 ) − T0 (S − S0 ) + m + mgz, (2.10)
2
and the exergy change of a closed system during a process would be the difference
between the final and initial exergies of the system

4X = X2 − X1
C22 − C12
= (U2 − U1 ) + p0 (V2 − V1 ) − T0 (S2 − S1 ) + m + mg(z2 − z1 ). (2.11)
2
Notes

• If work is done on the closed system, 4X would represent the minimum amount
of useful work that needs to be supplied in moving the system from state 1 to
state 2

• For stationary closed systems, the kinetic and potential energy terms drop out

• The exergy of a closed system is either positive or zero. It cannot be negative

Equation (2.10) may also be expressed on the basis of unit mass thus

C2
x = (u − u0 ) + p0 (v − v0 ) − T0 (s − s0 ) + + gz. (2.12)
2
Similarly, the exergy change in a closed system per unit mass is expressed

C22 − C12
4x = (u2 − u1 ) + p0 (v2 − v1 ) − T0 (s2 − s1 ) + + g(z2 − z1 ). (2.13)
2

EXAMPLE 6
A 200-m3 rigid tank contains compressed air at 1 MPa and 300 K. Determine how
much work can be obtained from this air if the environment conditions are 100 kPa

31
and 300 K. Treat air as an ideal gas and neglect the kinetic and potential energies.
SOLUTION
The air in the tank is the system and since no mass crosses the system boundary during
the process, this is a closed system, therefore we shall use Eq. (2.10) while neglecting
the KE and PE terms.

T = T0 = 300 K
pV (1 x 106 )(200)
m= = = 2323 kg
RT 287 x 300
There being no temperature difference between the system and the environment, there
wouldn’t be a change in the internal energy, therefore

U − U0 = 0 (2.14)

Further,
   
mRT mRT0 p0
p0 (V − V0 ) = p0 − = mRT0 −1
p p0 p
100 x 103
 
= 2323 x 287 x 300 − 1 = −180 M J
1 x 106
(2.15)

and
 
T p p
T0 (S − S0 ) = T0 mcp ln − mR ln = −mRT0 ln
T0 p0 p0
1 x 106
= −2323 x 287 x 300 x ln
100 x 103
= −460.54 M J. (2.16)

Therefore,

X = −180 − (−460.54) = 280.54 M J. (2.17)

This is the maximum useful work that can be obtained from the compressed air stored
in the tank in the specified environment.

2.5 Exergy of a Flow Stream


A flowing fluid has an additional form of energy over and above that contained in a
non-flow one. This flow work is essentially the boundary work done by the fluid on the
fluid downstream. The exergy associated with flow work is equivalent to the exergy
associated with the boundary work in excess of the work done against the atmospheric

32
air at p0 to displace it by a volume v. Thus, if the flow work is pv and the work done
against the atmosphere is p0 v, the exergy associated with the flow energy is expressed

xf low = pv − p0 v = (p − p0 )v (2.18)

The expression for the total flow exergy of a flow stream is determined by adding the
flow energy to the non-flow exergy (Eq. (2.13)) thus

xf lowstream = xnonf low + xf low


C2
= (u − u0 ) + p0 (v − v0 ) − T0 (s − s0 ) + + gz + (p − p0 )v
2
C2
= (u + pv) − (u0 + p0 v0 ) − T0 (s − s0 ) + + gz
2
C2
= (h − h0 ) − T0 (s − s0 ) + + gz. (2.19)
2
This is the flow exergy. The exergy change of a fluid stream as it undergoes a process
from state 1 to 2 is
C22 − C12
4xf lowstream = (h2 − h1 ) − T0 (s2 − s1 ) + + g(z2 − z1 ) (2.20)
2
NOTES

• For fluid streams with negligible kinetic and potential energies, the KE and PE
terms can drop out.

• For ideal gases, we have the following expressions for the enthalpy and entropy
changes

h − h0 = cp (T − T0 ),
T p
s − s0 = cp ln − R ln .
T0 p0

The change of flow availability for an ideal gas (neglecting KE and PE changes)
may then be expressed as
 
T2 p2
4xf lowstream = cp (T2 − T1 ) − T0 cp ln − R ln . (2.21)
T1 p1

• The exergy of a flow stream can (at pressures below environmental pressure) be
negative.

EXAMPLE 7
Refrigerant-134a is to be compressed from 0.14 MPa and -100 C to 0.8 MPa and 500 C

33
steadily by a compressor. Taking the environment conditions to be 200 C and 95 kPa,
determine the exergy change of the refrigerant during this process and the minimum
work input that needs to be supplied to the compressor per unit mass of refrigerant.
Neglect the kinetic and potential energies.
SOLUTION
The compressor is the system and since mass crosses the system boundary during the
compression process, this is a flow system.
Inlet state: p1 = 0.14 MPa, T1 = −100 C

h1 = 246.36 kJ/kg,
s1 = 0.9724 kJ/kg.K.

Exit state: p2 = 0.8 MPa, T2 = 500 C

h2 = 286.69 kJ/kg,
s2 = 0.9802 kJ/kg.K.

The exergy change is given by Eq. (2.20), while neglecting the KE and PE changes,
thus

4xf lowstream = (h2 − h1 ) − T0 (s2 − s1 )


= (286.69 − 246.36) kJ/kg − (293)(0.9802 − 0.9724) kJ/kg
= 38.04 kJ/kg. (2.22)

The refrigerant’s exergy increases by 38.04 kJ/kg during compression. This increase
in exergy (corresponding to the reversible work in that environment) represents the
minimum work that needs to be supplied to the compressor, thus

win,min = x2,f lowstream − x1,f lowstream


= 4xf lowstream
= 38.04 kJ/kg.

2.6 Second Law Efficiency


The second law efficiency (η2 ), also referred to as process effectiveness [6], is devel-
oped with the concept of availability. But first, we shall recap on some conventional
efficiencies.

34
2.6.1 Conventional Efficiencies
2.6.1.1 Thermal Efficiency
Some devices receive heat and convert it to work, e.g., the heat engine. The fraction
of the heat input that is converted to net work output is a measure of the performance
of the heat engine and is called thermal efficiency, ηth
net work output Wnet,out
ηth = = , (2.23)
total heat input Qin
and since Wnet,out = Qin − Qout , the thermal efficiency may also be expressed as
Qout
ηth = 1 − . (2.24)
Qin
In general, cyclic devices of practical interest like heat engines, refrigerators and heat
pumps operate between a high-temperature medium (or reservoir) at a temperature
TH , and a low-temperature medium (or reservoir) at temperature TL . We can then
define two heat quantities:

1. QH : The magnitude of heat transfer between the cyclic device and the high-
temperature medium at temperature TH .

2. QL : The magnitude of heat transfer between the cyclic device and the low-
temperature medium at temperature TL .

For the heat engine, we may then write

Wnet,out = QH − QL , (2.25)
Wnet,out
ηth =
QH
QL
=1− . (2.26)
QH
The thermal efficiency is a measure of how efficiently a heat engine converts the heat to
work, and is always less than unity. For example, an ordinary spark-ignition automobile
engine has a thermal efficiency of about 25%, while diesel engines could go upto 40%
[5].
2.6.1.2 Coefficient of Performance
The thermal efficiency described in the previous section cannot be greater than unity.
However the coefficient of performance, COP, which expresses the efficiency of refrig-
erators and heat pumps, may or may not be greater than 1 for refrigerators, and is
always greater than 1 for heat pumps. The analysis described next is also covered
under refrigeration in Section 1.8.

35
A refrigerator is a cyclic device that transfers heat from a low-temperature
medium to a high-temperature medium with the objective of maintaining the refriger-
ated space at a low temperature. Its efficiency is expressed in terms of coefficient of
performance (COPR ) as
desired output QL
COPR = = , (2.27)
required input Wnet,in
where QL is the heat removed from the refrigerated space. By the conservation of
energy principle for a cyclic device,

Wnet,in = QH − QL , (2.28)

therefore,
QL 1
COPR = = . (2.29)
QH − QL QH /QL − 1
A heat pump on the other hand is a cyclic device that also transfers heat from
a low-temperature medium to a high-temperature medium but with the objective of
maintaining a heated space at a high temperature, e.g., absorbing heat from cold
outside air and supplying it to a house during winter. Its efficiency is expressed in
terms of coefficient of performance (COPP ) defined thus
desired output QH
COPP = = , (2.30)
required input Wnet,in
or
QH 1
COPP = = . (2.31)
QH − QL 1 − QL /QH
Comparing Eq. (2.29) and Eq. (2.31) shows that

COPP = COPR + 1, (2.32)

implying that, while the efficiency of a refrigerator may or may not be greater than
unity, that of the heat pump is always greater than unity.

2.6.2 Determination of Second Law Efficiency


The conventional efficiencies described above are not a realistic measure of the per-
formance of engineering devices because they make no reference to the best possible
performance. This deficiency is overcome by way of the second law efficiency, de-
fined as the ratio of the actual thermal efficiency to the maximum possible (reversible)
thermal efficiency under the same conditions. Thus,
ηth
η2 = (heat engines). (2.33)
ηth,rev

36
The second law efficiency may be defined in general for work-producing devices (heat
engines, turbines etc) as
Wu
η2 = , (2.34)
Wrev
and such a definition applies to processes as well as to cycles. For work-consuming
devices, e.g., compressors and refrigerators, the second law efficiency is defined thus
Wrev
η2 = , (2.35)
Wu
and more specifically for refrigerators and heat pumps, it can be expressed in terms of
the coefficient of performance as
COP
η2 = . (2.36)
COPrev
We may generalize further and define the second law efficiency in terms of exergy, thus
exergy recovered exergy destroyed
η2 = =1− . (2.37)
exergy supplied exergy supplied
In a reversible process, the entire exergy supplied is recovered as there are no irre-
versibilities. If none of the supplied exergy is recovered, then the second law efficiency
is zero, but at all times, the sum of the exergy recovered and that destroyed (the
irreversibility) should equal the exergy supplied.
EXAMPLE 8
Consider two heat engines, both having a thermal efficiency of 30%. Engine A is sup-
plied with heat from a source at 600 K, while engine B is supplied heat from a source
at 1000 K. Both engines reject heat to a medium at 300 K. Determine the second law
efficiencies of the two engines. Discuss.
SOLUTION
The two engines can at best perform as reversible engines, in which case their efficien-
cies would be:
 
TL 300 K
ηrev,A = 1 − =1− = 50%,
TH A 600 K
 
TL 300 K
ηrev,B = 1 − =1− = 70%.
TH B 1000 K

Based on Eq. (2.33), the second law efficiencies of the two engines are:
0.3
η2,A = = 0.60,
0.5
0.3
η2,B = = 0.43.
0.7

37
Thus, though engine B has a greater work potential available to it than engine A (70%
of heat supplied, as compared to 50% for A), it converts only 43% of the available
work potential to useful work. Engine B on the other hand converts 60% of the work
potential available to it into useful work, and thus performs better than engine A.
Tutorial Problems 5

1. Steam enters an adiabatic turbine at 6 MPa, 6000 C, and 80 m/s and leaves at 50
kPa, 1000 C, and 140 m/s. If the power output of the turbine is 5 MW, determine
(a) The reversible power output [5.84 MW]
(b) The second law efficiency of the turbine [85.6%]
Assume the surroundings to be at 250 C.

2. Steam is throttled from 9 MPa and 5000 C to a pressure of 7 MPa. Determine


the decrease in exergy of the steam during this process [32.3 kJ/kg]
Assume the surroundings to be at 250 C.

3. Which has the capacity to produce more work in a closed system: 1 kg of steam
at 800 kPa and 1800 C, or 1 kg of R-134a at 800 kPa and 1800 C? [623 kJ, 5.0
kJ]
Take T0 = 250 C and p0 = 100 kPa.

4. A piston-cylinder initially contains 2 L of air at 100 kPa and 250 C. Air is now
compressed to a final state of 600 kPa and 1500 C. The useful work input is 1.2
kJ. Assuming the surroundings are at 100 kPa and 250 C, determine
(a) The exergy of the air at the initial and final states [0, 0.171 kJ]
The minimum work that must be supplied to accomplish this compression process
[0.171 kJ]
(c) The second law efficiency of this process [14.3%]

ASSIGNMENT 2 - Due on
Steam enters a turbine steadily at 3 MPa and 4500 C at a rate of 8 kg/s and exits
at 0.2 MPa and 1500 C. The steam loses heat to the surrounding air at 100 kPa and
250 C at a rate of 300 kW, and the kinetic and potential energy changes are negligible.
Determine:
(a) The actual power output
(b) The maximum possible power output
(c) The second law efficiency

38
3.0 The Ideal Gas and Gaseous Mixtures
3.1 The Perfect gas
At temperatures much higher than critical, fluid vapour tends to obey the equation,
pv
= constant = R (3.1)
T
In practice, no gas will obey this equation though many gases tend towards it. An
imaginary (ideal) gas that obeys the equation above perfectly is called a perfect gas.
Equation (3.1) then becomes the characteristic equation of state of a perfect gas. The
constant R is the specific gas constant and each gas has a distinct one [2].
Another form of equation (3.1) for a gas of mass m (kg) is,

pV = mRT (3.2)

(Note that V=mv ). In terms of molar mass M of the substance, also called molecular
weight,

pV = nM RT (3.3)
m
that is, M = , where n is the amount of substance or the number of moles of the
n
substance. The units of n are kmol, therefore the units of M are kg/kmol.
From equation (3.3), we can define,

M R = Ru (3.4)

where Ru is the molar gas constant (universal gas constant), such that

pV = nRu T or (3.5)
pV
= Ru (3.6)
nT
The value of Ru is 8314.5 J/kmol K. Equation (3.6) represents Avogadro’s hypothesis,
V
that is the same for all gases at the same value of p and T [2]. This hypothesis is
n
also referred to as Avogadro’s law and may alternatively be stated as:
Equal volumes of gases at the same temperature and pressure contain the same number
of particles [4].
When the molar mass of a gas is known, then its specific gas constant can be
calculated e.g. for air, the molar mass is 29 kg/kmol, therefore,
Ru 8314.5
R = = ≈ 287J/kgK (3.7)
M 29
Tutorial Problems 6

39
1. A vessel of volume 0.2 m3 contains nitrogen at 1.013 bar and 150 C. If 0.2 kg
of nitrogen is now pumped into the vessel, calculate the new pressure when the
vessel has returned to its initial temperature. The molar mass of nitrogen is 28
kg/kmol, and may be assumed to be a perfect gas. [1.87 bar]

2. A certain perfect gas of mass 0.01 kg occupies a volume of 0.003 m3 at a pressure


of 7 bar and a temperature of 1310 C. The gas is allowed to expand until the
pressure is 1 bar and the final volume is 0.02 m3 . Calculate

(a) the molar mass of the gas [16 kg/kmol]


(b) the final temperature [111.50 C]

3.1.1 Specific heat capacities


There are an infinite number of ways in which heat may be added between two temper-
atures in a gas, hence a gas could have an infinite number of specific heat capacities,
unlike for liquids and solids.
For reversible non-flow processes, two specific heat capacities are defined:

1. Specific heat capacity at constant pressure, cp ,

dQ = mcp dT (3.8)

For a perfect gas, cp is constant at all pressures and temperatures. Equation (3.8)
may be integrated to,

Q = mcp (T2 − T1 ) (3.9)

2. Specific heat capacity at constant volume, cv ,

dQ = mcv dT (3.10)

Again, the value of cv for a perfect gas is constant at all pressures and tempera-
tures. Therefore, integrating equation (3.10) gives,

Q = mcv (T2 − T1 ) (3.11)

3.1.2 Joule’s law


Joule’s law states that the internal energy of a perfect gas is a function of the absolute
temperature only, u = f (T ), and the function may be evaluated by considering a unit

40
mass of perfect gas heated at constant volume,

dq − dw = du
dq = du (dw = 0 when volume remains constant)
= cv dT
dq = du = cv dT , which integrates to
u = cv T + K (3.12)

For a perfect gas, it may be assumed that u = 0 at T = 0 such that the constant K is
zero. Therefore,

u = cv T (3.13)

For a mass m of perfect gas

U = mcv T (3.14)

Therefore, from equation (3.14), it can be deduced that the gain in internal energy
for a perfect gas between any two states, for any process (reversible or irreversible), is
given by,

U2 − U1 = mcv (T2 − T1 ) or 4U = mcv 4T (3.15)

3.1.3 Relationship between specific heat capacities


From the non-flow energy equation, Q − W = 4U . For a perfect gas, equation (3.15)
may thus be rewritten,

Q − W = mcv 4T (3.16)

For a constant pressure process,

W = p4V

Using equation (3.2), pV1 = mRT1 while pV2 = mRT2 , therefore

p4V = p(V2 − V1 ) = mR(T2 − T1 ) = mR4T

Therefore,

W = mR4T

Substituting in equation (3.16),

Q − mR4T = mcv 4T

41
Therefore,

Q = m(cv + R)4T (3.17)

For a constant pressure process, from equation (3.9),

Q = mcp 4T (3.18)

Equating equations (3.17) and (3.18),

m(cv + R)4T = mcp 4T

Therefore,

cv + R = cp (3.19)

3.1.4 Specific enthalpy of a perfect gas


Specific enthalpy is given by h = u + pv, for a perfect gas, pv = RT (equation (3.1))
and from Joule’s law (equation (3.13)), u = cv T . Therefore,

h = cv T + RT = (cv + R)T = cp T

For a mass m of a perfect gas

H = mcp T (3.20)

3.1.5 Ratio of specific heat capacities


cp
= γ (3.21)
cv
Since cp > cv (from equation (3.19)), then γ > 1. Some typical values are:

• γ = 1.6 for helium - a monoatomic gas (He)

• γ = 1.4 for oxygen - a diatomic gas (O2 )

• γ = 1.3 for carbon dioxide - a triatomic gas (CO2 )

• γ = 1.22 for ethane (C2 H6 )

From equation (3.19), other relationships can be derived:

cp − cv = R ,dividing through by cv ,
cp R
−1 =
cv cv

42
Substituting from equation (3.21),

R
γ−1 = , from which,
cv
R
cv = (3.22)
γ−1

multiplying equation (3.22) by γ,

γR
γcv = cp = (3.23)
γ−1

EXAMPLE 9
A perfect gas has specific heat capacities as follows:
cp = 0.846 kJ/kgK, cv = 0.657 kJ/kgK. Calculate the specific gas constant and the
molar mass of the gas. [0.189 kJ/kgK, M=44 kg/kmol]
Tutorial Problems 7

1. The relative molecular mass of CO2 is 44. In an experiment, the value of γ for
CO2 was found to be 1.3. Assuming that CO2 is a perfect gas, calculate the
specific gas constant R, and the specific heat capacities at constant pressure and
constant volume, cp and cv . [0.189 kJ/kgK; 0.819 kJ/kgK; 0.63 kJ/kgK]

2. Calculate the internal energy and enthalpy of 1 kg of air occupying 0.05 m3 at 20


bar. If the internal energy is increased by 120 kJ as the air is compressed to 50
bar, calculate the new volume occupied by 1 kg of the air. [250.1 kJ/kg; 350.1
kJ/kg; 0.0296 m3 ]

3. When a certain perfect gas is heated at constant pressure from 150 C to 950 C,
the heat required is 1136 kJ/kg. When the same gas is heated at constant vol-
ume between the same temperatures the heat required is 808 kJ/kg. Calculate
cp , cv , γ, R and the molar mass of the gas. [14.2 kJ/kgK; 10.1 kJ/kgK; 1.405; 4.1
kJ/kgK; 2.028 kg/kmol]

3.2 Gaseous mixtures


When perfect gases mix at a known pressure in a closed vessel of volume V and
temperature T, each of the gases contributes to the total pressure by an amount called
the partial pressure [2] (the partial pressure of each constituent gas is defined as that
pressure which the gas would exert if it occupied alone the whole volume occupied by
the mixture, at the same temperature). The relationship between the partial pressures
of the constituent gases is expressed by Dalton’s law as:

43
The pressure of a mixture of gases is equal to the sum of the partial pressures of the
constituent gases.
Therefore, if we had a mixture of gases: A, B, C, ... of masses mA , mB , mC , ... re-
spectively and partial pressures pA , pB , pC , ... respectively, then the total mass m of the
P
mixture is given by the law of conservation of mass as, m = mA +mB +mC +... = mi .
The total pressure p of the mixture is given by Dalton’s law as,
X
p = pA + pB + pC + ... = pi (3.24)

Air is the most common mixture and its analysis (both volumetric-by volume; and
gravimetric-by mass) are given in table 3.2 while table 3.3 shows the approximate
analysis in which air is deemed to be composed of oxygen and atmospheric nitrogen
[2].

Constituent Symbol Molar Mass(kg/kmol) Analysis


By Volume (%) By Mass (%)
Oxygen O2 31.999 20.95 23.14
Nitrogen N2 28.013 78.09 75.53
Argon Ar 39.948 0.93 1.28
Carbon dioxide CO2 44.010 0.03 0.05

Table 3.2: Volumetric and gravimetric analysis of air

Constituent Symbol Molar Mass(kg/kmol) Analysis


By Volume (%) By Mass (%)
Oxygen O2 32.0 21.0 23.3
Nitrogen N2 28.0 79.0 76.7
Table 3.3: Approximate analysis of air

EXAMPLE 10
A vessel of volume 0.4 m3 contains 0.45 kg of carbon monoxide and 1 kg of air at 150 C.
Calculate the partial pressure of each constituent, and the total pressure in the vessel.
Use the approximate gravimetric analysis of air and take the molar masses of CO, O2
and N2 as 28, 32 and 28 kg/kmol respectively [ pO2 = 0.436bar; pN2 = 1.64bar; pCO =
P
0.962bar; p = pi = 3.038bar]

44
3.2.1 Gibbs-Dalton’s law
Dalton’s law was reformulated to include a second statement that the thermodynamic
properties of a gas mixture are simply the sum of the properties of the constituent
gases. This led to the Gibbs-Dalton law:
The internal energy, enthalpy and entropy of a gaseous mixture are respectively equal
to the sum of the internal energies, enthalpies and entropies of the constituents. Each
constituent has that internal energy, enthalpy and entropy which it would have if it
occupied alone that volume occupied by the mixture at the temperature of the mixture
[2].
For internal energy, for example,
X
mu = mA uA + mB uB + mC uC + ... = mi ui (3.25)

3.2.2 Amagat’s (or Leduc’s) law


Consider a volume V of a gaseous mixture at temperature T, and pressure p, consisting
of constituents A, B and C. Let each of the constituents be compressed to pressure p,
equal to the total pressure of the mixture, while keeping the temperature constant.

Figure 3.1: Amagat’s law

Recall, pV = mRT , from Fig. 3.1 (a),

pA V
mA = (3.26)
RA T
and from Fig. 3.1 (b),

pVA
mA = (3.27)
RA T

45
Equating equations (3.26) and (3.27) gives,

pA V pVA
= , from which,
RA T RA T
pA
VA = V, or in general,
p
pi
Vi = V, and for all the constituents,
p
X X pi V X
Vi = V = pi (3.28)
p p
P
From equation (3.24), p = pi , therefore,
X
Vi = V, (3.29)

that is, the volume of a mixture of gases is equal to the sum of the volumes of the
individual constituents when each exists alone at the pressure and temperature of the
mixture. This is the law of partial volumes sometimes called Amagat’s or Leduc’s law
[2].
By extension (from Avogadro’s law), the total amount of substance in the vessel
must equal the sum of the amounts of substance of the individual constituents,
X
n = ni (3.30)

3.2.3 Molar mass and specific gas constant for a gas mixture
For any gas in a gas mixture of a total mass m occupying a total volume V at a
temperature T, it can be shown that:

1. The specific gas constant for the mixture is expressed


X mi
R = Ri (3.31)
m

2. The molar mass for the mixture is expressed:


X Vi
M = Mi , or, (3.32)
V
Xni
M = Mi , or, (3.33)
n
X pi
M = Mi (3.34)
p

46
3.2.4 Specific heat capacities for a gas mixture
3.2.4.1 Specific heat capacity at constant volume
P
From Gibbs-Dalton law, mu = mi ui ; and from Joule’s law for a perfect gas, u = cv T .
Hence for a gaseous mixture,
X X
mcv T = mi cvi T = T mi cvi
X
mcv = mi cvi
X mi
cv = cvi (3.35)
m
3.2.4.2 Specific heat capacity at constant pressure
P
From Gibbs-Dalton’s law, mh = mi hi ; and from section 3.1.4, the specific enthalpy
for a perfect gas is given by, h = cp T . Therefore, for a gaseous mixture,
X X
mcp T = mi cpi T = T mi cpi
X
mcp = mi cpi
X mi
cp = cpi (3.36)
m
3.2.4.3 Relationship between the specific heat capacities
From equations (3.35) and (3.36),
X mi X mi
cp − cv = cpi − cvi
m
Xm m
i
= (cpi − cvi ) (3.37)
m
and,

cpi − cvi = Ri , therefore,


X mi
cp − cv = Ri
m
X mi
Recall, R = Ri , therefore for a mixture, cp − cv = R, meaning that as well,
m
cp R γR
γ = , cv = , cp = (3.38)
cv γ−1 γ−1

Tutorial Problems 8

1. A mixture of 1 kmol of CO2 and 3.5 kmol of air is contained in a vessel at 1 bar
and 150 C. The volumetric analysis of air can be taken as 21% oxygen and 79%
nitrogen. Calculate for the mixture:

(a) the masses of CO2 , O2 and N2 , and the total mass [mCO2 = 44kg; mO2 =
23.55kg; mN2 = 77.5kg; m = 145.05kg]

47
(b) the molar mass and the specific gas constant for the mixture [M=32.5
kg/kmol; R= ]
(c) the specific volume of the mixture [v=0.7437 m3 /kg]

Take the molar masses of carbon, oxygen and nitrogen as 12 kg/kmol, 32 kg/kmol
and 28 kg/kmol respectively.

2. A vessel contains a gaseous mixture of composition by volume: 80% H2 and 20%


CO. It is desired that the mixture should be made in the proportion 50% H2 and
50%CO by removing some of the mixture and adding some CO. Calculate per
kilomole of mixture the mass of mixture to be removed, and the mass of CO to
be added. The pressure and temperature in the vessel remain constant during
the procedure.
Take the molar mass of hydrogen and carbon monoxide as 2 kg/kmol and 28
kg/kmol [2.7 kg of mixture removed; 10.5 kg of CO added]

3. The gas in an engine cylinder has a volumetric analysis of 12%CO, 11.5%O2 , and
76.5%N2 . The temperature at the beginning of expansion is 10000 C and the gas
mixture expands reversibly through a volume ratio of 7 to 1, according to a law
pv1.25 =constant. Calculate the work done and the heat flow per unit mass of gas.
The values of cp for the constituents are as follows:
cp for CO2 = 1.271 kJ/kgK
cp for O2 = 1.11 kJ/kgK
cp for N2 = 1.196 kJ/kgK
[w = 537.5 kJ/kg; heat supplied = 83.9 kJ/kg]

48
4.0 Psychrometry
4.1 Dry and Atmospheric Air
Air is a mixture of nitrogen, oxygen, and small amounts of other gases. Air in the
atmosphere normally contains some water vapour (moisture) and is referred to as
atmospheric air [5], while air without any moisture is called dry air. Normally
the amount of moisture in air is small, yet it is crucial to human comfort and hence
an important consideration in air-conditioning applications. The temperature of air in
air-conditioning applications ranges from about -10 to about 500 C. In this range, air is
treated as an ideal (perfect) gas with a constant cp = 1.005 kJ/kg.K. As for the moisture
in air: At 500 C, the saturation pressure of water is 12.3 kPa. At pressures below this,
water vapour can be treated as an ideal gas with negligible error, even when existing
as a saturated vapour. The water vapour then obeys the ideal-gas relation pv = RT .
The atmospheric air can then be treated as an ideal-gas mixture whose pressure is the
sum of the partial pressures of dry air (pa ) and that of water vapour (pv ), i.e.,

p = pa + pv . (4.1)

The partial pressure of water vapour, which is the pressure it would exert if it existed
alone at the temperature and volume of the atmospheric air, is normally called vapour
pressure, and since the water vapour is treated as an ideal gas, its enthalpy is a
function of temperature only, i.e., h = h(T ). Further, the enthalpy of water vapour in
air can be taken as equal to the enthalpy of saturated vapour at the same temperature,
i.e.,

hv (T, low p) ∼
= hg (T). (4.2)

In the range -10 to 500 C, the average value of cp for water vapour is taken to be 1.82
kJ/kg, and its enthalpy (in kJ/kg) is determined approximately from the equation

hg (T ) ∼
= 2500.9 + 1.82 T where T is in 0 C, (4.3)

where 2500.9 kJ/kg is the enthalpy of the water vapour at 00 C.


EXAMPLE 11
The enthalpy of water vapour at 400 C would be

2500.9 + 1.82(40) = 2573.7 kJ/kg,

which deviates by only 0.2 kJ/kg from the actual figure of hg read from the water
vapour tables (i.e., 2573.5 kJ/kg).

49
4.2 Specific and Relative Humidity of Air
Specific humidity (also absolute humidity or humidity ratio), denoted by ω, is defined
as
kg water vapour mv pv V /Rv T pv /Rv pv
ω= = = = = 0.622 , (4.4)
kg dry air ma pa V /Ra T pa /Ra pa
or
0.622pv
ω= , (4.5)
p − pv
where p is the total pressure. The more the water vapour in the air, the higher the
specific humidity, up until the air can hold no more moisture, at which point we refer
to the air as saturated air. Any moisture introduced into the saturated air will
condense. Moisture content in the air affects how comfortable we feel. The comfort
level largely depends on the actual amount of moisture the air holds (mv ) relative
to the maximum amount of moisture the air can hold at the same temperature (mg ).
The ratio of the two is the relative humidity, denoted φ, thus
mv pv V /Rv T pv
φ= = = , (4.6)
mg pg V /Rv T pg
where pg = psat at temperature, T . If we combine Eqs. (4.5) and (4.6), we have
0.622φpg
ω= and, (4.7a)
p − φpg

ωp
φ= . (4.7b)
(0.622 + ω)pg
Note that φ ranges between 0 and 1 and is dependent upon temperature.
The total enthalpy of atmospheric air (which is a mixture of dry air and water
vapour) is the sum of the respective enthalpies of the two,

H = Ha + Hv = ma ha + mv hv , (4.8)

and since mostly the amount of dry air in the mixture remains constant while that of
moisture varies, we may express the total enthalpy of atmospheric air per unit mass of
dry air by dividing through by ma ,

h = ha + ωhv , (4.9)

or

h = ha + ωhg , since hv ∼
= hg . (4.10)

50
The ordinary temperature of atmospheric air is referred to as dry-bulb temperature.
Other forms of temperature will be discussed in due course.
EXAMPLE 12
A 5-m x 5-m x 3-m room contains air at 250 C and 100 kPa at a relative humidity of
75%. Assuming the dry air and water vapour in the room to be ideal gases, determine:
(a) The partial pressure of dry air
(b) The specific humidity
(c) The masses of dry air and water vapour in the room
SOLUTION
For air, cp = 1.005 kJ/kg.K.
For water at 250 C, hg = 2546.6 kJ/kg from tables, or 2546.4 kJ/kg from Eq. (4.3) (we
can use this latter value).
(a)Partial pressure:

pv = φpg = φpsat,250 C = (0.75)(3.166) = 2.37 kP a,


pa = p − pv
= 100 − 2.37
= 97.63 kP a.

(b) Specific humidity:

0.622pv (0.622)(2.37)
ω= = = 0.0151 kg water vapour/kg dry air. (4.11)
p − pv 100 − 2.37

(c) Enthalpy

h = ha + ωhv ∼
= cp T + ωhg
= (1.005)(25) + (0.0151)(2546.4)
= 63.58 kJ/kg dry air. (4.12)

(d) Masses: The volume of each gas is equal to the volume of the room, while the
masses of the dry air and water vapour are determined from the ideal-gas relations

Va = Vv = Vroom = (5)(5)(3) = 75 m3 ,
pa Va (97.63)(75)
ma = = = 85.61 kg,
Ra T (0.287)(298)
pv V v (2.37)(75)
mv = = = 1.29 kg,
Rv T (0.4615)(298)

alternatively for the water vapour, mv = ωma = (0.0151)(85.61) = 1.29 kg.

51
4.3 Dew-point Temperature
When the excess moisture in the air condenses on a cool surface, the result is dew. The
dew-point temperature, Tdp is defined as the temperature at which condensation begins
when the air is cooled at constant pressure, i.e., Tdp is the saturation temperature of
water corresponding to the vapour pressure

Tdp = Tsat,pv . (4.13)

The ordinary and dew-point temperatures of saturated air are identical. When you
notice dew forming on your bottle of cold drink on a hot and humid day, this means that
the temperature of the drink is below the dew-point temperature of the surrounding
air.

4.4 Adiabatic Saturation and Wet-bulb Tempera-


tures
One way of determining the relative humidity is to determine the dew-point temper-
ature of air, after which the vapour pressure can be determined. This is simple but
impractical. An alternative is to use the adiabatic saturation process, see Fig. 4.1.
The system consists of a long insulated channel containing a pool of water. A steady
stream of unsaturated air with unknown specific humidity ω1 and a temperature T1 is
passed through the channel. As the air flows over the water, some water evaporates
and mixes with the airstream. The moisture content of air increases during this process
while the temperature decreases since part of the latent heat of vapourization of the
water that evaporates comes from the air. For a long-enough channel, the airstream
exits as saturated air, i.e., φ = 100% at temperature T2 , which is referred to as the
adiabatic saturation temperature.
If makeup water is supplied to the channel at the rate of evaporation at temperature
T2 , the adiabatic saturation process above can be analyzed as a steady-flow process. It
involves no heat or work interactions, and the KE and PE changes can be neglected.
Mass balance:

ṁa1 = ṁa2 = ṁa (mass flow rate of dry air remains constant),
ṁw1 + ṁf = ṁw2 (mass flow rate of vapour in the air increases by an amount
equal to the rate of evaporation ṁf ),
ṁa ω1 + ṁf = ṁa ω2 .

Thus,

ṁf = ṁa (ω2 − ω1 ).

52
Figure 4.1: Schematic and T-s diagram for the adiabatic saturation process of water

Energy balance:

Ėin = Ėout (since Q̇ = 0 and Ẇ = 0),


ṁa h1 + ṁf hf 2 = ṁa h2 ,

or

ṁa h1 + ṁa (ω2 − ω1 )hf 2 = ṁa h2 .

Dividing through by ṁa ,

h1 + (ω2 − ω1 )hf 2 = h2 ,

53
or

(cp T1 + ω1 hg1 ) + (ω2 − ω1 )hf 2 = (cp T2 + ω2 hg2 ),

to yield
cp (T2 − T1 ) + ω2 hf g2
ω1 = , (4.14)
hg1 − hf 2
where, from Eq. (4.7a)
0.622φ2 pg2
ω2 = ,
p2 − φ2 pg2
and since φ2 = 100%,
0.622pg2
ω2 = . (4.15)
p2 − pg2
Therefore, the specific humidity (and relative humidity) of air can be determined from
Eqs. (4.14) and (4.15) by measuring the pressure and temperature of air at the inlet
and exit of an adiabatic saturator.
If the bulb of an ordinary thermometer is covered in a cotton wick saturated with
water and air blown over the wick, the temperature determined by this configuration
is called the wet-bulb temperature. Though the basic principle is similar to that
of the adiabatic saturation, this is a more practical approach because in the earlier
method, a long channel or spray mechanism would normally be required to achieve
saturation conditions at the exit.
HOMEWORK 3 : Describe the principle of the wet-bulb thermometer.
What is a sling psychrometer?
EXAMPLE 13
The dry- and wet-bulb temperatures of atmospheric air at 1 atm (101.325 kPa) pres-
sure are measured with a sling psychrometer and determined to be 250 C and 150 C
respectively. Determine:
(a) The specific humidity
(b) The relative humidity
(c) The enthalpy of the air
SOLUTION
Saturation pressure of water is 1.7057 kPa at 150 C, and 3.1698 kPa at 250 C. At room
temperature, cp for air is 1.005 kJ/kg.K.
(a) Specific humidity
cp (T2 − T1 ) + ω2 hf g2
ω1 = ,
hg1 − hf 2

54
where T2 is the wet-bulb temperature and,

0.622pg2 (0.622)(1.7057)
ω2 = =
p2 − pg2 (101.325 − 1.7057)
= 0.01065 kg H2 O/kg dry air.

Thus,

1.005(15 − 25) + 0.01065(2465.4)


ω1 =
(2546.5 − 62.982)
= 0.00653 kg H2 O/kg dry air.

(b) Relative humidity

ω 1 p2 (0.00653)(101.325)
φ1 = =
(0.622 + ω1 )pg1 (0.622 + 0.00653)(3.1698)
= 0.332 or 33.2%.

(c) Enthalpy

h1 = ha1 + ω1 hv1 ∼
= cp T1 + ω1 hg1
= (1.005)(25) + (0.00653)(2546.5)
= 41.75 kJ/kg dry air.

4.5 The Psychrometric Chart


These are charts designed for use in air-conditioning applications to determine the state
of the air thereby saving on lengthy calculations. Figure 4.2 shows the basic features
of a psychrometric chart:

• On the horizontal axis are the dry-bulb temperatures

• The specific humidity is shown on the vertical axis

• On the left of the chart is a curve referred to as the saturation curve - all the
saturated air states are located on this curve. It is also the curve of 100% relative
humidity

• Lines of constant wet-bulb temperature are straight lines sloping downwards to


the right

• Lines of constant specific volume are also straight lines sloping downwards to the
right, but they are steeper than those of wet-bulb temperatures

55
• Lines of constant enthalpy lie very nearly parallel to the lines of constant wet-
bulb temperature (in some charts, the constant wet-bulb temperature lines are
used as constant-enthalpy lines)

Figure 4.2: Schematic diagram of a psychrometric chart

For saturated air, the dry-bulb, wet-bulb and dew-point temperatures are identical.
As such, the dew-point temperature of atmospheric air at any point on the chart can
be determined by drawing a horizontal line (a line of ω = constant or pv = constant)
from the point to the saturated curve. The temperature value at the intersection point
is the dew-point temperature.
Air conditioning processes are also suitably visualized using psychrometric charts.
An ordinary heating or cooling process appears as a horizontal line on the chart if no
humidification or dehumidification is involved (ω = constant). Note that any deviation
from a horizontal line indicates that moisture is added or removed from the air during
the process.
EXAMPLE 14
Consider a room that contains air at 1 atm, 350 C and 40% relative humidity. Using
the psychrometric chart, determine:
(a) The specific humidity
(b) The enthalpy

56
(c) The wet-bulb temperature
(d) The dew-point temperature
(e) The specific volume of the air
SOLUTION
At a given total pressure, the state of atmospheric air is completely specified by two
independent properties such as the dry-bulb temperature and the relative humidity.
Other properties are determined by directly reading their values at the specified state.
(a) Specific humidity: Draw a horizontal line from the specified state (T, ω) to the
right until it intersects with the ω axis. The value of ω is read off as 14.25/1000 =
0.01425 kg H2 O/kg dry air.
(b) Enthalpy: Draw a line parallel to the h = constant lines from the specified state
(T, ω) until it intersects the enthalpy scale, giving h = 71.5 kJ/kg dry air.
(c) Wet-bulb temperature: Draw a line parallel to the Twb = constant lines from the
specified state until it intersects the saturation line, giving Twb = 240 C.
(d) Dew-point temperature: Draw a horizontal line from the specified state to the left
until it intersects the saturation line, giving Tdp = 19.40 C.
(e) Specific volume: Consider the distances between the specified state and the v
= constant lines either side of this state. By interpolation, the specific volume is
determined to be v = 0.893 m3 /kg dry air.
Note that values read from the chart involve reading errors, they are not exact.

4.6 Human Comfort and Air Conditioning


It is important to condition a given space to give comfort to the occupants. This
involves heating/cooling, humidifying/dehumidifying, cleaning and even deodorizing
the air to the occupants’ desires. At any one time, the human body generates heat
depending on the activity. For an average adult male, the heat generated is about 87 W
when sleeping, 115 W when resting or doing office work, and upto 440 W when doing
heavy physical work. However, the inner temperature is maintained relatively constant
at about 370 C and as long as the body can dissipate the waste heat comfortably, the
body will feel comfortable in the given environment.
The body has a way of dealing with a diversity of environments. In a cold en-
vironment, the body tries to minimize heat loss by cutting down blood circulation
near the skin. This lowers the skin temperature and thus the heat transfer rate to the
surroundings. A low skin temperature however causes discomfort. Putting barriers
to heat transfer, e.g., by dressing heavily, reduces heat loss too. Cuddling, or putting
hands between the legs, may also help to reduce heat loss from the body, by reducing

57
the surface area of heat transfer. Alternatively, the amount of heat generated by the
body may be increased, say by way of exercising.
In hot environments, dressing lightly makes it easier for heat to get away from
our bodies. We may also reduce the amount of activity to minimize the rate of heat
generation in the body. Turning on the fan in an enclosed space continuously replaces
the warmer air layer forming around our bodies by the cooler air from the other parts
of the space.
In all, the comfort of the human body primarily depends on three factors:

• Dry-bulb temperature: This is the most important index of comfort. Most


people prefer a temperature of between 22 and 270 C.

• Relative humidity: This affects the amount of heat a body can dissipate
through evaporation (perspiration). Most people prefer 40 to 60% relative hu-
midity.

• Air motion: This removes the warm, moist air that builds up around the body
and replaces it with fresh air. Air motion thus improves heat rejection by both
convection and evaporation. It should be strong enough to remove heat and
moisture from the vicinity of the body, yet gentle enough to be unnoticed. Most
people prefer an airspeed of about 15 m/min.

Other factors that may affect comfort are air cleanliness and odour.

4.7 Air Conditioning Processes


Air conditioning processes are the processes that help to maintain a living space at
the desired temperature and humidity. These include simple heating, simple cooling,
humidifying, dehumidifying, or a combination of these. Most air conditioning processes
can be modelled as steady-flow processes, whose energy balance is expressed
X X
(Q̇in − Q̇out ) − (Ẇout − Ẇin ) = ṁh − ṁh. (4.16)
out in

4.7.1 Simple Heating and Cooling


In a simple heating or cooling process, the specific humidity remains constant (ω = con-
stant) since there is no humidification or dehumidification. A heating process proceeds
in the direction of increasing dry-bulb temperature and follows a line of constant spe-
cific humidity on the psychrometric chart, thus appearing as a horizontal line. However,
even though the specific humidity remains constant, the relative humidity decreases in
a heating process.

58
A cooling process is similar to the heating process only that the dry-bulb tem-
perature decreases while the relative humidity increases. The conservation of energy
equation for a heating or cooling process is simply

Q̇ = ṁa (h2 − h1 ), (4.17)

where h1 and h2 are the enthalpies per unit mass of dry air at the inlet and exit of the
heating/cooling section respectively, while ṁa is the mass flow rate of the air.

4.7.2 Heating with Humidification


Problems resulting from the low relative humidity associated with the simple heating
can be eliminated by humidifying the heated air. Once the air is passed over heating
coils, it is then passed over a humidifying section where steam could be used or water
could be sprayed into the airstream. When steam is used, there is additional heating,
such that T3 > T2 (Fig. 4.3(b)). Use of water results in cooling such that T3 < T2 .

(a) (b)

Figure 4.3: (a) Simple heating (b) Heating and dehumidifying

4.7.3 Cooling with Dehumidification


In the case of simple cooling (ω = constant), the relative humidity may increase to
undesirable levels, making it necessary to remove some moisture from the air by dehu-
midifying it. This requires cooling the air below its dew-point temperature (Fig. 4.4).
Hot moist air enters the cooling section at state 1. As it passes through the cooling
coils, its temperature decreases and its relative humidity increases at constat specific
humidity. If this cooling section is sufficiently long, air reaches its dew point (state
x). Further cooling beyond point x results in the condensation of part of the moisture
in the air. The air remains saturated during the entire condensation process (100%
relative humidity) till the final state (state 2) is reached. The condensate is removed

59
at state 2 after which the cool saturated air at this state is either routed directly to
the room, or heated to a more comfortable temperature (if it is too low) before being
routed to the room.

Figure 4.4: Cooling and dehumidification

The air-conditioning processes described above are summarized on the psychro-


metric chart of Fig. 4.5.

60
Figure 4.5: Air-conditioning processes

4.7.4 Evaporative Cooling


This is used as an alternative to conventional cooling that uses the refrigeration cycle,
because of the high initial and operating cost of refrigeration. In evaporative cooling,
as water evaporates, the latent heat of vapourization is absorbed from the water body
and the surrounding air, which are subsequently cooled in the process.
The evaporative cooling process is essentially identical to the adiabatic saturation
process, and follows a line of constant wet-bulb temperature on the psychrometric
chart.
HOMEWORK 4
Read and make brief notes about cooling towers.
Tutorial Problems 9

61
1. Air enters a 40-cm diameter cooling section at 1 atm, 320 C, and 30% relative
humidity at 18 m/s. Heat is removed from the air at a rate of 1200 kJ/min.
Determine:

(a) The exit temperature (24.40 C)


(b) The exit relative humidity of the air (46.6%)
(c) The exit velocity (17.6 m/s)

2. Air at 1 atm, 150 C, and 60% relative humidity is first heated to 200 C in a heating
section and then humidified by introducing water vapour. The air leaves the
humidifying section at 250 C and 65% relative humidity. Determine:

(a) The amount of steam added to the air (0.0065 kg H2 O/kg dry air)
(b) The amount of heat transfer to the air in the heating section (5.1 kJ/kg
dry air)

3. Atmospheric air at 1 atm, 300 C, and 80% relative humidity is cooled to 200 C
while the mixture pressure remains constant. Calculate:

(a) The amount of water, in kg/kg dry air, removed from the air (0.0069 kg
H2 O/kg dry air)
(b) The cooling requirement, in kg/kg dry air, when the liquid water leaves the
system at 220 C (27.3 kJ/kg dry air)

ASSIGNMENT 3 - Due on

1. The air in a room has a dry-bulb temperature of 220 C and a wet-bulb temperature
of 160 C. Assuming a pressure of 100 kPa, determine:

(a) The specific humidity


(b) The relative humidity
(c) The dew-point temperature

2. The air in a room has a pressure of 1 atm, a dry-bulb temperature of 240 C, and
a wet-bulb temperature of 170 C. Using the psychrometric chart, determine:

(a) The specific humidity


(b) The enthalpy, in kJ/kg dry air
(c) The relative humidity

62
(d) The dew-point temperature
(e) The specific volume of the air, in m3 /kg dry air

Figure 4.6: Psychrometric chart

63
4.7.5 Mixing of Airstreams
In some air conditioning applications, two airstreams require to be mixed, e.g., in large
buildings (like hospitals) where the conditioned air is mixed with a certain fraction of
fresh outside air before being routed into the interior of the building. This is done
by merging two streams of air. Usually, heat transfer with the surroundings is very
small and thus the mixing process is assumed to be adiabatic. Additionally, mixing
processes normally involve no work interactions, and the changes in KE and PE are
negligible. Consequently, the energy and mass balances for the adiabatic mixing of the
two airstreams reduce to

ṁa1 + ṁa2 = ṁa3 mass of dry air, (4.18)


ṁv1 + ṁv2 = ṁv3 mass of water vapour (4.19)
which is re-expressed through Eq. (4.4),
ω1 ṁa1 + ω2 ṁa2 = ω3 ṁa3 , (4.20)
ṁa1 h1 + ṁa2 h2 = ṁa3 h3 energy. (4.21)

If we eliminate ṁa3 from the above relations, we obtain


ṁa1 ω2 − ω3 h2 − h3
= = . (4.22)
ṁa2 ω3 − ω1 h3 − h1
Thus, when two airstreams at two different states (say 1 and 2 in Fig. 4.7) are mixed
adiabatically, the state of the mixture (state 3) lies on the straight line connecting
states 1 and 2 on the psychrometric chart, and the ratio of the distances 2-3 and 3-1
is equal to the ratio of mass flow rates ṁa1 and ṁa2 .
Tutorial Problem 10
Saturated air leaving the cooling section of an air-conditioning system at 140 C at a rate
of 50 m3 /min is mixed adiabatically with the outside at 320 C and 60% relative humidity
at a rate of 20 m3 /min. Assuming that the mixing process occurs at a pressure of 1
atm, determine for the mixture the:
(a) Specific humidity (0.0122 kg H2 O/kg dry air)
(b) Relative humidity (89%)
(c) Dry-bulb temperature (18.80 C)
(d) Volume flow rate (69.94 m3 /min)

64
Figure 4.7: Adiabatic mixing of two airstreams

65
REFERENCES
[1] Z. Warhaft, An Introduction to Thermal-Fluid Engineering (The Engine and the
Atmosphere). Cambridge University Press, 1997.

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


gists. Longman, 1993.

[3] M. D. Burghardt, Engineering Thermodynamics. Harper Collins, 1993.

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

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

[6] Li, Kam W., Applied Thermodynamics: Availability Method and Energy Conver-
sion. Taylor & Francis, 1995.

66

You might also like