You are on page 1of 222

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/363478207

Thermodynamic Analyses of Energy-Utilisation Systems Using


Excel

Book · September 2022

CITATIONS READS
0 219

1 author:

Mohamed El-Awad
Colleges of Applied Sciences, Sohar, Oman
66 PUBLICATIONS   150 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Use of Excel as a modelling platform for fluid-thermal analyses View project

Development of a Multi-Substance Add-In for Thermofluid Analyses using Excel View project

All content following this page was uploaded by Mohamed El-Awad on 04 October 2022.

The user has requested enhancement of the downloaded file.


Thermodynamic Analyses of Energy-
Utilisation Systems Using Excel

Mohamed M. El-Awad
2 Mohamed M. El-Awad
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 3

Thermodynamic Analyses of Energy-


Utilisation Systems Using Excel
4 Mohamed M. El-Awad
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 5

Thermodynamic Analyses of Energy-Utilisation


Systems Using Excel

Mohamed M. El-Awad
Associate Professor (resigned),
University of Technology and Applied Sciences
College of Applied Sciences, Suhar
University of Khartoum
Faculty of Engineering, Khartoum
6 Mohamed M. El-Awad

October, 2022

This book is dedicated to my students at

UTAS (CAS Suhar),

U. of K.,

UNITEN,

and UPM

The Thermax add-in can be found at:

https://www.academia.edu/76387165/ThermaX_A_Multi_Substance_Excel_Add_i
n_for_the_Analyses_of_Thermo_fluid_Systems
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 7

Preface

With the devastating effects of large-scale energy utilisation spreading worldwide like
fire, proper design and operation of energy-utilisation systems is becoming increasingly
critical. Since the design of these systems is mainly based on the principles of engineering
thermodynamics, understanding and properly applying these principles by engineering
students is needed more than ever before. In this respect, computers and computer-aided
methods can play an important role by making the application of these principles more
realistic and more accurate compared to traditional analytical methods that need many
idealisations and simplifying assumptions to be applied in order to model the systems.
Two examples of such simplifications are the air-standard assumptions used in the
analyses of internal-combustion engines and the constant specific-heat method in the
analyses of gas-turbine and similar cycles. Apart from enabling the analytical models to
be developed without such simplifications, computer-aided models are more useful for
performing parametric and design optimisation analyses of the system to minimise its
environmental impact and maximise its economic feasibility.

Thermodynamics is usually considered by junior engineering students to be a difficult


subject because of the numerous principles and laws it involves and because some of
these concepts are rather abstract and difficult to prove and apply. As for other
engineering courses, computer software and computer-aided methods can improve the
learning process by making the students more interested and by enabling them to deal
with difficult and time-consuming types of analyses. In this respect, this book shows how
Microsoft Excel can be used as an effective educational platform for thermodynamic
analyses by using suitable add-ins for fluid properties. Compared to the commercial
engineering software that is dedicated to such analyses, the main advantage of Excel for
this purpose is its wide availability on computers and mobile phones. From an educational
point of view, the Excel-based models allow the students to build white-box models from
the basic principles. By allowing the students to develop computer-aided models for their
analyses, Excel helps them to pay more attention to the principles pertinent to these
analyses and applications.

As a modelling platform for engineering analyses, Excel provides a rich library of


mathematical functions and powerful graphic tools for data visualisation. The Solver add-
in that comes with it enables the students to perform constrained and multi-variable
optimisation analyses of energy systems. Although Excel itself does not provide
functions for determining fluid properties which are necessary for thermodynamic
analyse, such functions can be developed as custom functions with the Visual Basic for
Applications (VBA) programming language that comes with Microsoft Applications. A
number of Excel add-ins that provide such functions have been developed by academic
and research institutions in the past. The present book uses the Thermax add-in that has
been developed by the author for educational purposes. The fluids covered by Thermax
include ideal gases, water and superheated steam, synthetic and natural refrigerants for
8 Mohamed M. El-Awad

vapour-compression refrigeration systems, humid air for psychrometric analyses, two


aqua solutions for vapour-absorption refrigeration, and atmospheric air at various
temperatures. The thermodynamic analyses presented in this book confirm the adequacy
of the Excel-Thermax platform as an educational tool for analysing various types of
power generation and refrigeration cycles, air-conditioning processes, and combustion
processes. Moreover, Excel’s two iterative tools Goal Seek and Solver enable the students
to deal with iterative solutions and time-consuming parametric and optimisation studies.

By adopting a problem-based learning approach, the book is intended to supplement


rather than substitute traditional teaching methods. The Excel-based modelling platform
is used to deal with different types of thermodynamic analyses that range from the
analyses of the basic gas power-generation and vapour-compression refrigeration cycles
to the analyses of vapour-absorption refrigeration cycles. Within this range, the topics
covered include the analyses and optimisation of multi-stage compression and cascade
VCR refrigeration systems, air-conditioning systems, combustion analyses, and single-
effect and two-effect vapour-absorption systems. Most of the cases considered for
analyses in the book are adopted from popular textbooks and published work so that the
results obtained can be verified. While relevant exercises are provided at the end of the
first four chapters, more challenging exercises are given as an appendix to the book.
Bearing in mind that property add-ins developed by other academic and research
institutions can be used to extend the range of thermodynamic analyses introduced in the
book, it is hoped that the book and the Excel-based modelling platform can also
encourage student research relevant to energy-utilisation systems.

Acknowledgements
The development of Thermax would not have been possible without benefitting from the
efforts of many other people who have made the fruits of their efforts available in the
open literature or the internet. In this respect, my special gratitude goes to the Mechanical
Engineering Department at the University of Alabama (USA) whose initiatives both
inspired and helped me throughout the writing of this book. My gratitude and
appreciation also go to the College of Applied Sciences in Suhar (Oman) who provided
the needed financial support for my work and helped me to prepare the material covered
in this book. I am also indebted the unfailing support received from my colleagues and
students in Malaysia, Sudan, and Oman, and hope that they find this book a fair repay.
Last, but not least, my thanks and gratitude go to my beloved family for their tolerance
and unfailing support desperately needed to complete the work.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 9

CONTENTS

Nomenclature x

1. Introduction 1
1.1. Thermodynamic analyses of energy-utilisation systems 2
1.2. Advantages of computer-aided thermodynamic analyses 7
1.3. Excel add-ins for fluid properties 9
1.4. Closure 10
References 11

2. Thermax property functions 13


2.1. The name style for Thermax property functions 14
2.2. Property functions provided by Thermax 16
2.2.1. Functions for ideal gases (Gas) 16
2.2.2. Functions for water and superheated steam (Wat) 17
2.2.3. Functions for vapour-compression refrigerants (Ref) 18
2.2.4. Functions for psychrometric analyses (Psy) 19
2.2.5. Functions for absorption refrigeration solutions (LiB and NH3) 20
2.2.6. Functions for combustion and chemical reactions (Chm) 22
2.2.7. Functions for air at standard atmospheric pressure (Air) 22
2.3. Using Thermax property functions in Excel formulae 23
2.3.1. Accessing Thermax functions via the Function Wizard 23
2.3.2. Direct use of Thermax functions in Excel formulae 26
2.4. Analyses with variable specific heats using Thermax functions 29
2.5. Closure 30
References 31
Exercises 33

3. Analyses of gas power cycles 35


3.1. The ideal Brayton cycle 36
3.2. The regenerative Brayton cycle 40
3.3. Energy analysis of the Otto cycle 44
3.4. Exergy analysis of the Otto cycle 47
3.5. Closure 50
References 51
Exercises 51

4. Analyses of the Rankine cycle and the combined Brayton-Rankine cycle 53


4.1. The ideal simple Rankine cycle 54
4.2. The Rankine cycle with superheat and reheat 58
4.3. The regenerative Rankine cycle with a single feedwater heater 62
4.4. The regenerative Rankine cycle with two feedwater heaters 68
4.5. The combined Brayton-Rankine cycle 67
10 Mohamed M. El-Awad

4.5.1. First-law analysis of the combined cycle 72


4.5.2. Second-law analysis of the combined cycle 75
4.6. Closure 79
4.7. References 79
4.8. Exercises 80

5. Analyses of simple and multi-stage compression VCR cycles 83


5.1. The ideal vapour-compression refrigeration cycle 84
5.2. The actual vapour-compression refrigeration cycle 87
5.3. Optimisation analysis of the ideal two-stage compression cycle 90
5.3.1. The analytical model 91
5.3.2. The optimisation analysis 91
5.4. Optimisation analysis of the ideal three-stage compression cycle 94
5.4.1. The analytical model 95
5.4.2. The optimisation analysis 96
5.5. Closure 99
References 99
Exercises 99

6. Analyses of cascade VCR systems 101


6.1. The analytical model for the cascade refrigeration cycle 102
6.2. Optimisation of the ideal cascade refrigeration cycle using R134a 105
6.3. Analysis of the actual cascade cycle with R744-refrigerant pairs 108
6.4. Analysis of the cascade cycle with Refrigerants R507A and R23 111
6.5. Closure 113
References 114

7. Analyses of air-conditioning systems 115


7.1. The psychrometric chart and the basic psychrometric processes 116
7.2. Cooling with dehumidification 117
7.3. Adiabatic mixing of two air streams 121
7.4. The evaporative cooler 122
7.5. Wet cooling towers 125
7.6. Design analysis of an air-conditioning system 129
7.7. Closure 134
References 135

8. Development of a general Excel sheet for combustion analyses 137


8.1. The basic equations for combustion analyses 138
8.1.1. The combustion equation for a hydrocarbon fuel in air 138
8.1.2. The dew-point temperature of combustion products 140
8.1.3. Condensation of water vapour in the combustion products 142
8.1.4. The first-law of thermodynamics for steady-flow combustion 143
8.2. Development of a general Excel sheet for combustion analyses 144
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 11

8.3. Using the general sheet for combustion analyses of various fuels 146
8.4. Determining the adiabatic flame temperature with Goal-Seek 151
8.5. Dealing with unsupported fuels 152
8.6. Using macros in combustion analyses 154
8.7. Closure 156
References 156

9. Analyses of vapour-absorption refrigeration cycles 157


9.1. Analytical model for the single-effect VAR system 158
9.1.1. The analytical model 159
9.1.2. Analysis of ammonia-water VAR cycle 160
9.1.3. Analysis of water-lithium bromide VAR cycle 162
9.2. Analysis of a double-effect VAR system 165
9.3. Closure 169
References 170

Appendices 171
A. Ideal-gases, refrigerants and chemical reactants supported by Thermax 172
B. Thermodynamic properties of ideal gases and humid air 177
C. Verification of Thermax functions for superheated refrigerants 180
D. Additional first-law analyses of combustion processes 186
E. Second-law analyses of combustion processes 193
F. Additional exercises and projects 200

Subject index 208


12 Mohamed M. El-Awad

Nomenclature

Property Units
Air-fuel ratio -
Concentration (of an absorption solution) %
Density Kg/m3
o
Dry-bulb temperature C
Diffusivity m2/s
o
Dew-point temperature C
Dynamic viscosity μPa.s
Enthalpy (specific) kJ/kg
Enthalpy of formation (molar) kJ/kmol
Entropy (absolute molar) kJ/kmol.K
Enthalpy (molar) kJ/kmol
Entropy (specific) kJ/kg.K
Entropy of saturated liquid solution kJ/kg.K
Entropy of sturated vapour soliution kJ/kg.K
Entropy change with temperature (ideal gas) kJ/kg.K
Entropy change with temperature (molar) kJ/kmol.K
Gibbs function of formation (molar) kJ/kmol
Heat of vaporization of fuel kJ/kg
Humidity (specific or absolute) -
Heating value of fuel (higher and lower) kJ/kg.K
Humidity (relative) %
Internal energy (specific) kJ/kg
Internal energy (molar) kJ/kmol
Kinematic viscosity m2/s
Prandtl number -
Pressure kPa
Quality (dryness fraction) -
Relative pressure -
Refrigerant pressure kPa
Relative specific volume -
o
Refrigerant temperature C
Saturation pressure kPa
o
Saturation temperature C
Specific volume m 3/kg
Specific heat at constant pressure kJ/kg.K
Specific heat at constant pressure (molar) kJ/kmol.K
Surface tension mN/m
o
Temperature C
o
Temperature (wet-bulb) C
Temperature (absolute) K
Thermal conductivity mW/(m.oC)
o -1
Volumetric expansion coefficient C
Velocity of sound m/s
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 13

Chapter 1

Introduction
14 Mohamed M. El-Awad

The principles of engineering thermodynamics help us to determine the amount and


effectiveness of the energy transfer between an energy-utilisation system and its
surroundings in the form of work or heat. Thermodynamic analyses of the cycles that
involve phase changes of the working fluid, like the Rankine cycle, require the
determination of fluid properties at different phases of the fluid. For these cycles,
computer-aided methods are more convenient than traditional methods that use property
tables and charts. For gas cycles that do not involve phase changes computer-aided
methods are more accurate because they allow the application of the variable specific-
heat method of analysis instead of the approximate constant specific-heat method. This
chapter illustrates the application of thermodynamic principles for the analyses of energy-
utilisation systems and highlights the advantages of computer-aided methods for such
analyses. The chapter also introduces the Thermax add-in used in this book for
thermodynamic analyses of energy-utilisation systems with Microsoft Excel.

1.1. Thermodynamic analyses of energy-utilisation systems


The two basic thermodynamic principles are the conservation of mass principle and the
conservation of energy principle (the first-law of thermodynamics). The third principle,
which is the second-law of thermodynamics, accounts for the various losses that usually
occur in any energy-utilisation process. These three thermodynamic principles take
different mathematical expressions depending on whether the system is open or closed
and whether the flow steady or unsteady, compressible or incompressible. Consider the
air-compression system shown in Figure 1.1 that has two stages of compression separated
by an intercooler. Air enters the system at a temperature T1 and pressure P1. The first-
stage compressor, C1, compresses the air adiabatically to state 2, after which it enters the
intercooler where its temperature is reduced to T3. The second-stage compressor, C2, then
increases the air pressure to P4 at which the temperature increases to T4. Figure 1.2 shows
the compression process on a temperature-entropy diagram.

4
Air
C1 C2
1

2 Intercooler 3
Cooling
water

Figure 1.1. Schematic diagram of a two-stage air compressor with inter-stage


intercooling

The required compression work is divided between the two compressor stages depending
on their compression ratios and there is a certain value of the intermediate pressure (Pi)
that minimises the total work. The principles of thermodynamics help us to determine
this optimum intermediate pressure.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 15

T
P4
P2 = P3
4 2
P1

T1 3 1
Intercooling

s
Figure 1.2. T-s diagram of a two-stage air compressor with inter-stage intercooling

Treating the two compressor stages as steady-flow processes, and neglecting changes in
kinetic and potential energy, the first-law of thermodynamics leads to [1]:

q  w  hout  hin  (1.1)

Where q and w are the amounts of heat transfer and work transfer per unit mass flow of
air, respectively, and (hout –hin) is the resulting enthalpy change in the stage. Equation
(1.1) adopts the usual sign convention that heat into the system is positive, while work
into the system is negative.

Assuming the compression processes in both stages to be adiabatic (q=0) and reversible
means that the processes are isentropic as shown in Figure 1.2. Using an average specific
heat for air at constant pressure (cp), the compression work per unit mass flow of air in
stage 1 (w1) and in stage 2 (w2) can be determined from Equation (1.1) as follows:

w1  h2  h1   c p T2  T1  (1.2)

w2  h4  h3   c p T4  T3  (1.3)

Therefore, the total compression work in both stages (wtotal) is given by:

wtotal  w1  w2  c p T2  T1   T4  T3  (1.4)

Assuming perfect intercooling, i.e., T3 = T1, and rearranging Equation (1.4):

 T   T4   T   T4 
wtotal  c p T1 1  2   1    c p T1 2   2     (1.5)
 T1   T3    T1   T3 
16 Mohamed M. El-Awad

Since we already assumed the two compression processes to be isentropic and the specific
heat cp for air to be constant, the temperature ratios in Equation (1.5) can be converted
into pressure ratios by using the following relationships:

k 1
T2  P2  k
  (1.6)
T1  P1 
k 1
T4  P4  k
  (1.7)
T3  P3 

Where k is the ratio of specific heats (k=cp/cv); cv is the specific heat for air at constant
volume. Making another assumption that there is no pressure loss in the intercooler, then
P3 = P2= Pi. Substituting from Equations (1.6) and (1.7), Equation (1.5) becomes:

 k 1 k 1

  Pi  k  P4  k 
wtotal  c p T1 2        (1.8)
  P1   pi  

To see how the total compression work varies with the intermediate pressure Pi, let us
consider the specific case in which T1= 300K, P1=100 kPa, and P4 = 900 kPa. Using
Equation (1.8), the total compression work in the system was calculated for different
values of Pi and the result is shown in Figure 1.3. The figure shows that the value of Pi
at which the total compression work is minimal is around 300 kPa. Increasing or
decreasing Pi from this value will increase the compression work.

310
Total compression work (kJ)

300

290

280

270

260
0 200 400 600 800 1000
Intermediate pressure (kPa)
Figure 1.3. Variation of the total compression work with the intermediate pressure

The principles of thermodynamics are particularly useful for performance evaluation and
optimisation of power-generation and refrigeration systems. For example, consider the
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 17

regenerative steam-turbine power plant shown in Figure 1.4. This plant consists of a
boiler house for producing superheated steam, a high-pressure steam turbine (HPT), a
low-pressure steam turbine (LPT), a condenser, an open feed-water heater (FWH) and
two feed-water pumps. A fraction of the steam (y) is extracted after the HPT for
preheating the feed-water before going back to the boiler house. The extracted steam
reduces the work output from plant, but it also reduces the amount of heat added in the
boiler and its net effect is to increase the thermal efficiency of the plant. There is also a
certain extraction pressure for the steam at which the plant’s thermal efficiency attains a
maximum value. As shown below, the principles of thermodynamics can be used to
determine this optimum steam-extraction pressure.

1 2

HPT LPT
Boiler
House
y
3
1-y

7 FWH
Condenser
5
6 4
Pump 2 Pump 1
Figure 1.4. Schematic diagram of a regenerative steam-turbine power plant

The total specific work output from the two turbines (wout) and the total work input to the
two pumps (win) are given by:

wout  wHPT  wLPT (1.9)

win  wP1  wP 2 (1.10)

Where wHPT and wLPT are the specific work output from the high-pressure turbine and the
low-pressure turbine, respectively, and wP1 and wP2 are the specific work input in pump
1 and pump 2, respectively. Assuming the two turbines and the two pumps to be adiabatic
and neglecting the changes in kinetic and potential energies, the work output or input for
each device can be determined from the enthalpy difference across the device. Per each
kg of steam generated in the boiler, these are given by:

wHPT  h1  h2  (1.11)


18 Mohamed M. El-Awad

wLPT  1  y h2  h3  (1.12)

wP1  1  y h5  h4  (1.13)

w P 2  h7  h6  (1.14)

Mass and energy balance over the open feed-water heater gives:

yh2  1  y h5  1  h6 (1.15)

The net specific work output from the plant (wnet) is then given by:

wnet  wout  win (1.16)

The specific heat input to the boiler (qin) is determined by the relevant enthalpy change
as follows:

q in  h1  h7  (1.17)

Therefore, the thermal efficiency of the plant (η) can be calculated from:

  wnet / q in (1.18)

Both wnet and η depend on the fraction of steam extracted for regeneration (y); which in
turn depends on the extraction pressure (P2). Figure 1.5 shows the variation of y and η
with P2 for an ideal cycle in which P1 = 15 MPa, T1 = 600oC, and P4 = 10 kPa. The figure
shows that the cycle’s efficiency attains a maximum value of 45.55% when P2 is in the
range of 1000 kPa. It should be mentioned that the working fluid in the above power plant
changes phase from subcooled liquid water to superheated steam in the boiler, to
saturated mixture of water and steam in the low-pressure turbine, and returns to subcooled
water in the condenser. Therefore, appropriate property relationships, tables, or charts are
needed in order to determine the working fluid properties at different states.

The principles of thermodynamics are also needed for the analyses of air-conditioning
systems and processes and the analyses of the processes that involve combustion and
other chemical reactions. For such analyses, thermodynamics provides the basic
relationships needed to quantify the effects of fluid mixing and chemical reactions on the
properties of the working fluids and on the transfer of energy and effluents to or from the
fluid-thermal system.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 19

0.456 0.3

Fraction of extracted steam (y)


Thermal efficiency (η)
0.455 0.25

0.454 0.2

0.453 0.15

0.452 0.1
Efficiency
0.451 0.05
y
0.45 0
0 1000 2000 3000 4000
Extraction pressure (kPa)

Figure 1.5. The effect of intermediate pressure (P2) on the fraction of extracted steam
(y) and thermal efficiency (η) of a regenerative steam-turbine power plant

1.2. Advantages of computer-aided thermodynamic analyses


Apart from saving time and effort and eliminating possible human errors, computer-aided
thermodynamic of analyses offer a number of advantages over traditional method of
analyses that use property tables and charts. An important advantage of computer-aided
methods with respect to thermodynamic analyses is their ability to give more realistic
results by avoiding unnecessary simplification of the models and by using more accurate
estimations of fluid properties. Moreover, they offer reliable techniques for iterative
solutions and optimisation analyses. In what follows, these advantages are illustrated by
means of relevant examples.

A. Avoiding excessive simplification of the model


In many situations, traditional analytical methods adopt excessive simplifications of the
analytical models; which makes their results grossly deviate from the behaviour of real
systems. A good example of this situation is given by the models of internal-combustion
(IC) engines. Traditional air-standard models of IC engines, such as the Otto cycle and
the Diesel cycle, involve many simplifications such as neglecting heat-transfer and
friction losses, treating the combustion process as heat-addition from an external source,
and using constant specific heats. These assumptions enable the engine processes to be
represented by simple closed-form relations for calculating the amount of heat added to
the engine and net work from the engine. However, air-standard models usually
overestimate the engine’s output and thermal efficiency. By comparison, computer-aided
models of IC engines such as those described by Ferguson [2] take into consideration the
geometrical as well as the thermodynamic characteristics of the engines. These models,
which closely mimic the behaviour of actual IC engines, can be used to investigate the
effect of important design and operation factors such the ignition or injection timing on
the engine performance or the effect of engine speed on the specific fuel consumption.
20 Mohamed M. El-Awad

However, the formulation of these models leads to a set of ordinary differential equations
that need to be solved simultaneously by using a numerical solver such as the Newton-
Raphson method.

B. Accurate representation of fluid properties and processes


The behaviour of real gasses and vapours is frequently modelled by using the following
ideal-gas law:

Pv~  Ru T (1.19)

~
Where P is the absolute pressure of the gas, v is the molar specific volume, Ru is the
universal gas constant, and T is the absolute temperature of the gas. The ideal gas law
can be used with reasonable accuracy for determining the specific volume of a
superheated vapour, but when the temperature approaches the saturation line, the value
of the specific volume determined by the ideal-gas law departs significantly from the
actual volume. More accurate estimates can be obtained by using the Soave-Redlich-
Kwong (SRK) equation of state [1]:

RT a
P~u ~ ~ (1.20)
v  b v ( v  b)

Where the constants a, b and  are fluid-dependent. Figure 1.6 shows the deviations from
the tabulated values by those obtained from the ideal-gas law and the SRK equation of
state for refrigerant R134a at 0.2 MPa.

7
Error in specific volume (%)

6 Ideal-gas model
SRK model
5
4
3
2
1
0
250 300 350 400
Temperature K
Figure 1.6. Errors in the specific volume of R134a by the ideal-gas law
and the SRK equation of state
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 21

The figure shows that the error of the ideal-gas law is more that 2% even at high
temperatures and increases as the temperature approaches the saturation value, but the
accuracy of the SRK equation remained higher than 99% even close to the saturation line.
~
However, since the SRK equation is implicit in v , it cannot be used directly to determine
the specific volume, but has to be solved iteratively. A number of standard iterative
procedures (e.g. Newton-Raphson method) can be used to solve the equation, but they
are not convenient for hand calculations.

C. Dealing with iterative solutions and optimisation analyses


Thermodynamic analyses involving iterative solutions and optimisation analyses suit
computer-aided methods more than manual methods even for simple systems. A good
example of thermodynamic analyses that require iterative solutions is the determination
of the adiabatic flame temperature by a first-law analysis of the combustion process.
Optimisation analyses are needed for determining the best design for a certain energy-
utilisation system such as the optimum intermediate pressure for an air-compression
system, the optimum steam-extraction pressure for a regenerative Rankine cycle, and the
optimum intermediate pressures for multi-stage and cascade refrigeration systems. While
simple optimisation analyses that involve a single design parameter can be performed by
means of calculus techniques and graphic tools, optimisation analyses of complex
systems that involve multiple variables require the use of computer-aided techniques.

1.3. Excel add-ins for fluid properties


Considering its rich library of built-in functions, the simplicity of its user-interface, and
the flexibility of its graphical tools, Microsoft Excel is being increasingly used as a
teaching aid in various engineering subjects. Although Excel is mostly used for dealing
with simple computer-aided operations like matrix inversion and matrix multiplications,
it is equipped with other tools that make it a capable modelling platform for more
challenging types of “What-if” and optimisation analyses [3-5]. Two of these tools are
the Goal Seek command and the Solver add-in. The Visual Basic for Applications (VBA)
programming language that comes with Microsoft Office can be used for developing
customised user-defined functions (UDFs) and add-ins for thermodynamic analyses. The
Developer ribbon of Excel also allows the use of macros that remove the tedium of
parametric studies and repetitive calculations.

The main limitation of Excel as modelling platform for thermodynamic analyses is the
lack of built-in functions for fluid properties, but this problem could be solved by
developing suitable add-ins. The Thermotable add-in developed by the Mechanical
Engineering Department at the University of Alabama determines the thermodynamic
properties of ideal gases, water and superheated steam, and four refrigerants R134a, R22,
R410A, and R407C [6-9]. Goodwin [10] also developed an educational Excel add-in,
called TPX (Thermodynamic Properties for Excel), that determines the properties of five
gases (H2O, H2, O2, N2, CH4) and refrigerant R-134a. A number of property add-ins have
also been developed for research and industrial applications [11, 12]. The American
National Institute of Standards and Technology (NIST) developed REFPROP that
22 Mohamed M. El-Awad

provides thermophysical properties of various refrigerants and their mixtures [13]. An


open-source alternative to REFPROP, called CoolProp, was developed by Bell [14] at
the University of Liege. Optimized Thermal Systems also developed a commercial
alternative add-in to REFPROP called XProps [15].

This book presents an educational add-in, called Thermax, that enables Excel to be used
for thermodynamic analyses of various types of energy-utilisation systems [16]. Thermax
provides property functions for seven groups of working fluids used in energy-utilisation
systems that include:

- Twenty-nine ideal gases,


- Water and superheated steam,
- Twenty-eight refrigerants for vapour-compression refrigeration (VCR) systems,
- Two aqua solutions for vapour-absorption refrigeration (VAR) systems,
- Humid air for psychrometric analyses,
- Various types of fuels and chemically-reacting substances,
- Air at standard atmospheric pressure but various temperatures.

In addition to its seven groups of fluid-property functions, Thermax provides a small


group of user-defined functions that are useful for general thermofluid analyses. These
include two data interpolation functions, a Newton-Raphson solver for non-linear
equations, and a function that determines the diameters of standard-size pipes [17-19].

1.3. Closure
Chapter 2 describes the Thermax add-in and its fluid property functions in more details
and shows how these functions can be used in Excel’s formulae for thermodynamic
analyses. The four chapters that follow illustrate the use of Thermax functions for the
analyses of power-generation and refrigeration systems. Chapters 3 and 4 that deal with
the analyses of power-generation cycles give examples of analysing the Brayton cycle,
the Otto cycle, the Rankine cycles with and without regeneration, and the combined
Brayton-Rankine cycle. Chapter 5 deals with the analyses of single-stage and multi-stage
VCR cycles while Chapter 6 deals with the analysis and thermodynamic optimisation of
cascade VCR cycles.

The last three chapters, 7, 8, and 9, show how Thermax can be used to deal with
thermodynamic analyses in which the working fluids are mixtures of gaseous or liquid
substances. Chapter 7 deals with the analyses of psychrometric processes and illustrates
the use of Thermax property functions for design analyses of air-conditioning systems.
Chapter 8 focuses on combustion analyses and describes the development of a general
Excel sheet for steady-flow combustion analyses. Appendices D and E extend the general
Excel sheet to deal with combustion in closed systems and second-law analyses of
combustion processes. Chapter 9 illustrates the use of Thermax functions for the analyses
of single-effect and double-effect VAR cycles using ammonia-water and Lithium-
bromide-water solutions.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 23

References
[1] Y.A. Cengel, and M.A. Boles, Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[2] C.R. Ferguson, Internal Combustion Engines, John Wiley & Sons, 1986.
[3] A. Karimi, Using Excel for the thermodynamic analyses of air-standard cycles and
combustion processes, ASME 2009 Lake Buena Vista, Florida, USA.
[4] Z. Ahmadi-Brooghani, Using Spreadsheets as a Computational Tool in Teaching
Mechanical Engineering, Proceedings of the 10th WSEAS International Conference
on computers, Vouliagmeni, Athens, Greece, July 1315, 2006, 305-310
[5] S.A. Oke, Spreadsheet Applications in Engineering Education: A Review, Int. J.
Engng Ed. Vol. 20, 2004, No. 6, 893-901
[6] The University of Alabama, Mechanical Engineering, Excel for Mechanical
Engineering project, Internet: http://www.me.ua.edu/excelinme/index.htm (Last
accessed July 11, 2019).
[7] J. Huguet, K. Woodbury, R. Taylor, Development of Excel add-in modules for use
in thermodynamics curriculum: steam and ideal gas properties, American Society
for Engineering Education, 2008, AC 2008-1751.
[8] K. Mahan, J. Huguet, K. Woodbury, R. Taylor, Excel in ME: Extending and
refining ubiquitous software tools, American Society for Engineering Education,
2009, AC 2009-2295.
[9] L. Caretto, D. McDaniel, T. Mincer. Spreadsheet calculations of thermodynamic
properties, Proceedings of the 2005 American Society for Engineering Education
Annual Conference & Exposition, American Society for Engineering Education,
2005.
[10] D. G. Goodwin, "TPX: thermodynamic properties for Excel",
http://termodinamicaparaiq.blogspot.com/p/tpx_12.html (Last accessed July 11,
2019).
[11] T. K. Jack. Computerised calculations of thermo-physical steam and air properties,
Int. J. Pure Appl. Sci. Technol., 9(2) (2012), pp. 84-93 (Available online at
http://ijopaasat.in/yahoo_site_admin/assets/docs/3_IJPAST-276-
V9N2.139104407.pdf (Last accessed July 11, 2019).
[12] C.O.C. Oko and E.O.Diemuodeke. MS Excel spreadsheet add-in for
thermodynamic properties and process simulation of R152a, Energy Science and
Technology, Vol. 5, No. 2, 2013, pp. 63-69.
[13] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Reference Fluid
Thermodynamic and Transport Properties— REFPROP Version 8.0, User’s
Guide, National Institute of Standards and Technology, Physical and Chemical
Properties Division, Boulder, Colorado 80305, 2007.
[14] I. Bell, CoolProp. Available at https://sourceforge.net/projects/coolprop/files/
CoolProp/ (Last accessed July 11, 2019).
[15] Optimized Thermal Systems, XProps, A Quick and easy way to call refrigerant
properties, http://optimizedthermalsystems.com/index.php/products/xprops, (Last
accessed July 11, 2019).
24 Mohamed M. El-Awad

[16] M.M. El-Awad, Thermax: A Multi-Substance Excel Add-in for the Analyses of
Thermo-fluid Systems-
https://www.academia.edu/76387165/ThermaX_A_Multi_Substance_Excel_Add
_in_for_the_Analyses_of_Thermo_fluid_Systems
[17] M.M. El-Awad, Computer-Aided Thermofluid Analyses using Excel -
https://www.academia.edu/40015718/Computer_Aided_Thermofluid_Analyses_
using_Excel
[18] M.M. El-Awad, Optimisation Analyses of Fluid-Thermal Systems using Excel -
https://www.academia.edu/36307413/Optimisation_Analyses_of_Fluid_Thermal
_Systems_using_Excel
[19] M.M. El-Awad and M.S. Al-Saidi, Excel as an Educational Platform for Design
Analyses of Fluid-Thermal Systems, World Journal of Engineering and
Technology Vol.10 No.2, 434-443, doi: 10.4236/wjet.2022.102025.Scientific
Research Publishing.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 25

Chapter 2

Thermax property functions


26 Mohamed M. El-Awad

Thermax provides Excel with more than 100 custom functions for determining the key
thermo-physical properties of various working fluids used in energy-utilisation systems.
To help the user select the suitable function, the add-in functions have been divided into
seven groups such that each group deals with similar substances, i.e., ideal gases, water-
and-steam, vapour-compression refrigerants (VCRs), vapour-absorption refrigerants
(VARs), humid-air for psychrometric analyses, fuels-and chemical reactants for
combustion analyses, and air at standard pressure for general thermo-fluid analyses. Each
group of functions has a common indicative prefix; e.g. “Gas” for ideal gases, “Wat” for
water, and “Ref” for VCRs. In addition to its group, the name of each function shows its
output and input parameters. This chapter explains this name style with examples and
shows how it can be utilised by Excel’s function wizard to help the user find the required
function without having to memorise the names of all the functions. The chapter also lists
the functions provided by each group and describes the procedure for installing the add-
in and using its functions in Excel’s formulae for thermodynamic analyses.

2.1. The name style for Thermax property functions


Consider the two Thermax functions shown in Figure 2.1. The figure shows that the
names of these two property functions consist of the following three parts:

1. The first part is the function’s group, viz., “Gas” for ideal gases, “Wat” for water,
“Ref” for refrigerants, “Psy” for psychrometric analyses, “LiB” and “Nh3” for
the two binary solutions used in vapour-absorption refrigeration, “Chm” for
fuels and chemically reacting substances, and “Air” for air at atmospheric
pressure.
2. The second part is the function’s output property followed by an underscore, e.g.,
“h_” for enthalpy, “s_” for entropy, and “v_” for the specific volume.
3. The third part is the function’s input parameter(s), e.g., “PT” for pressure and
temperature (in oC) and “Ps” for pressure and entropy.

(a)

(b)
Figure 2.1. Examples of Thermax functions
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 27

The function Wath_Px shown in Figure 2.1.a determines the enthalpy of saturated water
at a pressure of 500 kPa and quality of 0.8. The first three letters in the name of this
function (Wat) refer to its group, which is water, immediately followed by the function’s
output, which is enthalpy (h). An underscore ( _ ) precedes the function’s two input
arguments, which are the pressure (P) and quality (x). Similarly, the name of the function
Refs_PT shown in Figure 2.1.b tells that it belongs to the refrigerants group (Ref) and
that it determines the entropy (s) of a superheated refrigerant from its pressure (P) and
temperature (T). This function requires three input parameters which are: (i) the
refrigerant name (“R134a”), (ii) the pressure (200 kPa), and (iii) the temperature (50oC).
Table 2.1 gives more examples for the add-in functions with their intended usage.

Table 2.1. Examples of Thermax property functions from different groups


# Thermax function Output
1 Gash_TK(“Air”,350) Determines the enthalpy (h) for air at 350K
Determines the entropy (s) of saturated water at
2 Wats_Tx(150,0.5)
a temperature of 150oC and quality of 0.5
Determines the enthalpy of superheated steam at
3 Wath_PT(90,150)
90 kPa and 150oC
Determines the saturation pressure (Psat) for
4 RefPsat_T(“R134a”,-5)
refrigerant R134a at -5oC.
PsyRh_PTSh(101,30, Determines relative humidity (ϕ) of humid air at
5 0.001) 101 kPa, 30oC, and specific humidity (ω) of
0.001 kg/kg
Determines the dynamic viscosity of air at
6 Airdv_T(25)
standard atmospheric pressure and 25oC.
7 Chm_f1("C6H6","M") Returns the molar mass for benzene (C6H6)
Determines the enthalpy of water-lithium
8 LiBh_TX(20,80)
bromide solution at 20oC and 80% concentration.

The following general points should be noted regarding Thermax property functions:

1. Three groups involve more than one substance; which are the Gas-group, the
Ref-group, and the Chm-group. The functions of these groups require the name
of the substance as the first input parameter.
2. In the Gas-group, the absolute temperature is represented by “TK” while in the
other groups the temperature in oC is represented by “T”.
3. The input and output properties in most of the functions are represented by one
or two letters, e.g. Gash_TK and GasTK_h, but there are few exceptions to this
rule. For example, some properties are represented by three letters like the
saturation pressure and saturation temperature, which are represented by Psat and
Tsat in the Wat and Ref groups. Also, the density of atmospheric air in the Air-
group is represented by Airrho.
28 Mohamed M. El-Awad

2.2. Property functions provided by Thermax


Thermax supports 29 ideal gases, water and superheated steam, 28 vapour-compression
refrigerants, two binary solutions used in vapour-absorption refrigeration, various fuels
and chemically reacting substances, humid air, and atmospheric air at various
temperatures. The functions provided by each group are described below.

2.2.1. Functions for ideal gases (Gas)


This group of Thermax functions determine the ideal-gas properties for the 29 gasses
listed in Appendix A. Based on the ideal-gas relationships given in Appendix A, the
group provides the 12 functions listed in Table 2.2.

Table 2.2. Functions for thermodynamic properties of ideal gases


# Function Input [Unit] Output [Unit]
1 GasM Gas name Molar mass, M [kg/kmol]
2 Gascp_TK Gas name, T[K] Specific heat, cp [kJ/kg.K]
3 Gash_TK Gas name, T[K] Specific enthalpy, h [kJ/kg]
4 Gasu_TK Gas name, T[K] Specific internal energy, u [kJ/kg]
5 Gass0_TK Gas name, T[K] Entropy component, s0 [kJ/kg.K]
6 GasPr_TK Gas name, T[K] Relative pressure, Pr [-]
7 Gasvr_TK Gas name, T[K] Relative volume, vr [K]
8 GasTK_h Gas name, h[kJ/kg] Absolute temperature, T [K]
9 GasTK_u Gas name, u[kJ/kg] Absolute temperature, T [K]
10 GasTK_s0 Gas name, s0[kJ/kg.K] Absolute temperature, T [K]
11 GasTK_Pr Gas name, Pr [-] Absolute temperature, T [K]
12 GasTK_vr Gas name, vr [K] Absolute temperature, T [K]

The first five functions in Table 2.2 determine the molar mass (M), specific heat (cp),
enthalpy (h), internal energy (u), and the part of entropy change due to temperature
change (s0), respectively. The two functions that follow, GasPr_TK and Gasvr_TK,
determine the relative pressure (Pr) and relative volume (vr) of the ideal gases,
respectively. The last five functions in the table are inversion functions that determine
the temperature of the ideal gas from its enthalpy (h), internal energy (u), temperature-
dependent entropy change (s0), relative pressure (Pr) or relative specific volume (vr).

With the exception of the first function, GasM, all the functions in this group involve the
gas absolute temperature as input or output. Note that the temperature in the names of all
the functions in this group is represented by “TK”, while in the other add-in groups the
temperature is represented by “T” only. An auxiliary function named “Gas_data”, not
shown in the table, stores the values of the four coefficients a0, a1, a2, and a3 in Equation
(A.1) for the 29 gases [1]. The same function is needed by the third function, Chm_f3,
in the combustion and chemical reactions group to determine the properties of the gases
involved in a combustion or chemical reaction process.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 29

2.2.2. Functions for water and superheated steam (Wat)


The functions in this group that determine the thermo-physical properties of water and
superheated steam are divided into two subgroups: (a) functions for the properties of
saturated water/steam mixture and (b) functions for the properties of superheated steam.
Properties of subcooled liquid water can be taken as the corresponding values of saturated
liquid water at the given temperature.

a) Properties of saturated water/steam mixture


The functions in this subgroup do not use mathematical formulae, but store and
interpolate the tabulated data for saturated water provided by ASHRAE [2]. Table 2.3
shows the input and output parameters of the 10 property functions that determine the
properties of saturated water/steam mixtures at a given temperature, in oC, with their
relevant units. The first function “WatPsat_T” returns the saturation pressure at a given
temperature. Properties of water as saturated liquid or saturated vapour can be obtained
from the functions number 2 to 9 by assigning the value of the quality x as 0 or 1,
respectively. The corresponding 10 custom functions that provide properties of saturated
water-steam mixture at a given pressure, in kPa, are listed in Table 2.4.

Table 2.3. Property functions for saturated water/steam at a given temperature


# Function Input [Unit] Output [Unit]
1 WatPsat_T T [oC] Saturation pressure, ps [kPa]
o
2 Watv_Tx T [ C], x [-] Specific volume, v [m3/kg]
o
3 Watu_Tx T [ C], x [-] Specific internal energy u [kJ/kg]
o
4 Wath_Tx T [ C], x [-] Specific enthalpy, h [kJ/kg]
5 Wats_Tx T [oC], x [-] Specific entropy, s [kJ/kg.K]
o
6 Watcp_Tx T [ C], x [-] Specific heat cp , [kJ/(kg·K)]
7 Watk_Tx T [oC], x [-] Thermal cond., k [mW/(m·K)]
8 Watdv_Tx o
T [ C], x [-] Viscosity,  [Pa·s]
9 Watvs_Tx T [oC], x [-] Velocity of sound, [m/s]
o
10 Watst_T T [ C] Surface tension, [mN/m]

Table 2.4. Property functions for saturated water/steam at a given pressure


# Function Input [Unit] Output/Unit
1 WatTsat_P P [kPa] Saturation temperature, Ts [oC]
2 Watv_Px P [kPa], x [-] Specific volume, v [m3/kg]
3 Watu_Px P [kPa], x [-] Specific internal energy u [kJ/kg]
4 Wath_Px P [kPa], x [-] Specific enthalpy, h [kJ/kg]
5 Wats_Px P [kPa], x [-] Specific entropy, s [kJ/kg.K]
6 Watcp_Px P [kPa], x [-] Specific heat cp , [kJ/(kg·K)]
7 Watk_Px P [kPa], x [-] Thermal cond., k [mW/(m·K)]
8 Watdv_Px P [kPa], x [-] Viscosity,  [Pa·s]
9 Watvs_Px P [kPa], x [-] Velocity of sound, [m/s]
10 Watst_P P [kPa] Surface tension, [mN/m]
30 Mohamed M. El-Awad

b) Properties of superheated steam


There are 8 functions in this subgroup as shown in Table 2.5. The first four functions
determine the specific volume (v), internal energy (u), enthalpy (h), and entropy (s) of
superheated steam given the pressure and temperature according to the formulae provided
by Irvine and Liley [3]. The remaining four functions are inverse functions that use
iteration to determine the temperature, enthalpy, or entropy from the pressure and another
property.

Table 2.5. Properties of superheated steam given the pressure and another property
# Function Input [Unit] Output [Unit]
1 Watv_PT P, T[oC] Specific volume, v [m3/kg]
2 Watu_PT P, T[oC] Specific internal energy u [kJ/kg]
3 Wath_PT P, T[oC] Specific enthalpy, h [kJ/kg]
4 Wats_PT P, T[oC] Specific entropy, s [kJ/kg.K]
5 WatT_Ph P, h[kJ/kg] Temperature, T [0C]
6 WatT_Ps P, s[kJ/kg.K] Temperature, T [0C]
7 Wath_Ps P, s[kJ/kg.K] Specific enthalpy, h [kJ/kg]
8 Wats_Ph P, h[kJ/kg] Specific entropy, s [kJ/kg.K]

2.2.3. Functions for vapour-compression refrigerants (Ref)


This group of functions provides the thermo-physical properties for the 28 refrigerants
listed in Appendix A, Table A2. As for water, the functions for saturated refrigerants do
not use mathematical formulae but store and interpolate the tabulated data provided by
ASHRAE [2] by using a linear interpolation function. Table 2.6 lists 11 functions in this
subgroup that determine the thermo-physical properties of saturated refrigerants at a
given pressure. Three of the 28 refrigerants are zeotropic blends, which are R404A,
R407C, R410A. For these three refrigerants, the saturation temperature means the bubble
temperature, while the dew temperature is given by the last function shown on Table 2.6,
RefTdew_P. Table 2.7 lists 10 functions that determine the refrigerants properties at a
given temperature.

Table 2.6. Properties of saturated refrigerants at a given pressure in kPa and quality
# Function Input Output Output unit
1 RefTsat_P Refrigerant name, P, x Saturation temperature, Tsat [0C]
2 Refv_Px Refrigerant name, P, x Specific volume, v [m3/kg]
3 Refu_Px Refrigerant name, P, x Specific internal energy, u [kJ/kg]
4 Refh_Px Refrigerant name, P, x Specific enthalpy, h [kJ/kg]
5 Refs_Px Refrigerant name, P, x Specific entropy, s [kJ/kg.K]
6 Refcp_Px Refrigerant name, P, x Specific heat, cp [kJ/kg.K]
7 Refvs_Px Refrigerant name, P, x Velocity of sound [m/s]
8 Refdv_Px Refrigerant name, P, x Dynamic viscosity, µ [µPa·s]
9 Refk_Px Refrigerant name, P, x Thermal conductivity, k [mW/(m·K)]
10 Refst_P Refrigerant name, P Surface tension [mN/m]
11 RefTdew_P Refrigerant name, P Dew temperature [0C]
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 31

Table 2.7. Properties of saturated refrigerants at a given temperature in oC and quality


# Function Input Output Output unit
1 RefPsat_T Refrigerant name, T, x Saturation temperature, Tsat [0C]
2 Refv_Tx Refrigerant name, T, x Specific volume, v [m3/kg]
3 Refu_Tx Refrigerant name, T, x Specific internal energy, u [kJ/kg]
4 Refh_Tx Refrigerant name, T, x Specific enthalpy, h [kJ/kg]
5 Refs_Tx Refrigerant name, T, x Specific entropy, s [kJ/kg.K]
6 Refcp_Tx Refrigerant name, T, x Specific heat, cp [kJ/kg.K]
7 Refvs_Tx Refrigerant name, T, x Velocity of sound [m/s]
8 Refdv_Tx Refrigerant name, T, x Dynamic viscosity, µ [µPa·s]
9 Refk_Tx Refrigerant name, T, x Thermal conductivity, k [mW/(m·K)]
10 Refst_T Refrigerant name, T Surface tension [mN/m]

Unlike the functions for saturated refrigerants, the thermodynamic properties of


superheated refrigerants are determined by using mathematical formulae and not by data
interpolation. The relevant formulae are given in Appendix B. Table 2.9 lists the 8
functions that apply these formulae to deal with superheated vapours of refrigerants. The
first 7 functions in Table 2.9 use Equations (B.7) to (B.13) to determine property values,
while the last function is an inversion function that uses an iterative solution. It should be
noted that these functions apply only for subcritical pressures. Appendix E verifies the
functions for R134a by comparing their estimations with the data given by ASHRAE [2].

Table 2.9. Properties of superheated refrigerants given the pressure in kPa


# Function Input [unit] Output [unit]
1 Refv_PT Refrigerant name, P, T [oC] v [m3/kg]
2 Refu_PT Refrigerant name, P, T [oC] u [kJ/kg]
3 Refh_PT Refrigerant name, P, T [oC] h [kJ/kg]
o
4 Refs_PT Refrigerant name, P, T [ C] s [kJ/kg.K]
5 RefT_Ph Refrigerant name, P, h [kJ/kg] T [oC]
6 RefT_Ps Refrigerant name, P, s [kJ/kg.K] T [oC]
7 Refh_Ps Refrigerant name, P, s [kJ/kg.K] h [kJ/kg]
8 Refs_Ph Refrigerant name, P, h [kJ/kg] s [kJ/kg.K]

2.2.4. Functions for psychrometric analyses (Psy)


The functions of this group provide the thermodynamic properties of moist atmospheric
air, which is a mixture of dry air and water-vapour. Appendix B shows the relations
commonly used in psychrometric analyses that are used for the development of these
functions. Table 2.10 lists 14 functions that are included in this group together with their
input and output arguments. The letters in the function names have the following
meanings: h (specific enthalpy), Db (dry bulb), Dp (dew point), P (pressure), Rh (relative
humidity), Sh (specific humidity), v (specific volume), Wb (wet bulb). Case 5 in Table
32 Mohamed M. El-Awad

2.1 shows how the function PsyRh_PTSh is used to determine the relative humidity for
air given its pressure, temperature, and specific humidity.

Table 2.10. Functions for psychrometric analyses


# Function Input [unit] Output [unit]
1 Psyv_PDbRh P [kPa], Tdb [oC], ϕ [%] v [m3/kg]
2 PsyRh_PDbSh P [kPa], Tdb [oC], ω[kg/kg] ϕ [%]
3 PsyRh_PDbWb P [kPa], Tdb [oC], Twb [oC] ϕ [%]
4 PsySh_PDbRh P [kPa], T [ C], ϕ [%]
o
ω [kg/kg]
5 PsySh_PDbWb P [kPa], Tdb [oC], Twb [oC] ω [kg/kg]
6 Psyh_PDbSh P [kPa], T [ C], ω[kg/kg]
o
h [kJ/kg]
7 Psyh_PDbRh P [kPa], T [oC], ϕ [%] h [kJ/kg]
8 PsyDp_PDbRh P [kPa], Tdb [ C], ϕ [%]
o
Tdp [oC]
o o
9 PsyDp_PDbWb P [kPa], Tdb [ C], Twb [ C] Tdp [oC]
10 PsyDb_PRhSh P [kPa], ϕ [%], ω[kg/kg] Tdb [oC]
11 PsyDb_PhSh P [kPa], h [kJ/kg], ω[kg/kg] Tdb [oC]
12 PsyWb_PDbRh P [kPa], Tdb [ C], ϕ [%]
o
Twb [oC]
13 PsyWb_PDbSh P [kPa], Tdb [oC], ω[kg/kg] Twb [oC]
14 PsyWb_PRhSh P [kPa], ϕ [%], ω[kg/kg] Twb [oC]

2.2.5. Functions for absorption refrigeration solutions (LiB and NH3)


Thermax provides two groups of property functions for the analyses of VAR systems: (a)
water-lithium bromide (H2O-LiBr) solution and (b) ammonia-water (NH3-H2O) solution.
The two groups use different formulae to determine properties of these solutions.

A. The H2O-LiBr solution (LiB)


To determine the enthalpy (h), refrigerant temperature (Tr), and refrigerant concentration
(X) of the H2O-LiBr solution, the functions use the formulae given by ASHRAE [2] and
to determine the entropy (s) and specific volume (v) the functions use the formulae given
by Patek and Klomfar [4] and by Patterson and Perez-Blanco [5], respectively. This group
needs functions that determine the saturation pressure and temperature of the refrigerant,
which is water in this case. These properties can be obtained by using the relevant
functions in the water group, but to make the group self-sufficient its functions determine
these properties from the Antoine equations, Equations (B.15) and (B.16). Table 2.11
lists the functions in the Lib-group.

B. The ammonia-water solution (NH3)


To determine the enthalpy of the ammonia-water solution, this group applies the formulae
given by Patek and Klomfar [6] and to determine the specific volume it applies that given
by Sun [7]. The entropy of NH3-H2O solution in the liquid state (sl) and the vapour state
(sv) are obtained by using the formulae given by El-Shaarawi et al [8]. The relations
between refrigerant pressure and temperature of an ammonia-water solution with
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 33

concentration are taken from Bourseau and Bugarel [9]. Table 2.12 lists the functions in
the NH3-group.

Table 2.11. Functions for water-lithium bromide solution


# Function Input [unit] Output [unit]
o
1 LiBh_TX T [ C], X [%] h [kJ/kg]
2 LiBv_TX T [oC], X [%] v [m3/kg]
o
3 LiBs_TX T [ C], X [%] s [kJ/kg.K]
o
4 LiBT_trX Tr [ C], X [%] T [oC]
5 LiBT_prX Pr [kPa], X [%] T [oC]
6 LiBT_hX h [kJ/kg], X [%] T [oC]
o
7 LiBTr_TX T [ C], X [%] Tr [oC]
8 LiBPr_TX T [oC], X [%] Pr [kPa]
9 LiBX_Ttr T [oC], Tr [oC] X [%]
o
10 LiBX_Tpr T [ C], Pr [kPa] X [%]
11 LiBT_sX s [kJ/kg.K], X [%] T [oC]
12 LiBpr_tr T [oC] Psat [kPa]
13 LiBtr_tr P [kPa] Tsat [oC]

Table 2.12. Functions for ammonia-water solution


# Function Input [unit] Output [unit]
o
1 NH3h_TX T [ C], X [%] h [kJ/kg]
2 NH3v_TX T [oC], X [%] v [m3/kg]
3 NH3sl_prX Pr [kPa], X [%] sl [kJ/kg.K]
4 NH3sv_prX Pr [kPa], X [%] Sv [kJ/kg.K]
5 NH3T_trX Tr [oC], X [%] T [oC]
6 NH3T_prX Tr [oC], X [%] T [oC]
7 NH3T_hX h [kJ/kg], X [%] T [oC]
8 NH3tr_TX T [oC], X [%] Tr [oC]
o
9 NH3pr_TX T [ C], X [%] Pr [kPa]
o o
10 NH3X_Ttr T [ C], Tr [ C] X [%]
11 NH3X_Tpr T [oC], Pr [kPa] X [%]
12 NH3pr_tr T [oC] Psat [kPa]
13 NH3tr_pr P [kPa] Tsat [oC]

The NH3-group also needs functions that determine the saturation pressure and
temperature of the refrigerant, which is ammonia. In order to make the group self-
sufficient, the group has its own functions that determine these properties instead of using
the functions provided by the refrigerants group. Accordingly, the saturation pressure
(Psat) of ammonia is obtained from the following equation [7]:
6
Psat  103  a i T i (2.1)
i 0
34 Mohamed M. El-Awad

Where, T is in oC, P in kPa and values of the six coefficients a1 to a6 can be found in Sun
[7]. The saturation temperature (Tsat) is obtained from Equation (2.1) by iteration.

2.2.6. Functions for combustion and chemical reactions (Chm)


The analyses of chemical reactions, e.g. combustion, involve mass and energy balances
of the substances involved as reactants or products that can be solids, liquids, or gases.
The gases involved are treated as ideal and, therefore, the group of functions provided by
the add-in for ideal gases can be used here. However, combustion analyses require
additional data like the heating value, enthalpy of formation, or Gibbs function. The
group provided by Thermax for the analyses of combustion processes and chemical
reactions consists of the three general functions shown in Table 2.13.

Table 2.13. Functions for thermodynamic properties of reacting systems


# Function Input argument Output property
1 Chm_f1 Substance; property M, h f0 , g 0f , s 0
2 Chm_f2 Substance; property M, QH, QL, AF, Hv, α, β, γ, δ
3 Chm_f3 Substance; property, temperature cp , h , u , s0

All three functions require as input arguments the substance name and the name of the
required property. In addition to that the third function requires the temperature of the
substance involved to be provided. The first function, Chm_f1, provides the molar mass
(M), molar enthalpy of formation ( h f0 ), molar Gibbs function of formation ( g 0f ), and
molar absolute entropy ( s 0 ) for the 29 reacting substances and reaction products that are
listed in Appendix A, Table A.3. Case 7 in Table 2.1 illustrates the use of this function.

The second function, Chm_f2, provides additional data needed for combustion analyses
involving the 20 fuels and reacting substances listed in Table A.4. These include values
of the molar mass (M), gravimetric higher and lower heating values (QH and QL),
stoichiometric air-fuel ratio (AFs) and heat of vaporisation (Hv). This function also
provides the carbon, hydrogen, oxygen, and nitrogen contents of the substance (α, β, γ,
δ) based on the general chemical representation CαHβOγNδ. The third function, Chm_f3,
determines four properties of ideal gases on a unit-mole basis.

2.2.7. Functions for air at standard atmospheric pressure (Air)


This group provides the thermo-physical properties required by fluid-dynamics and heat-
transfer analyses for air at standard atmospheric pressure. Based on the tabulated data
given by Cengel and Ghajar [10], the functions use a linear interpolation function to
determine the air density (ρ), specific heat (cp), thermal conductivity (k), thermal
diffusivity (α), dynamic viscosity (μ), kinematic viscosity (ν) and Prandtl number (Pr) at
temperatures in the range -150oC to 2000oC. Table 2.14 shows the names of the seven
functions in this group with their corresponding output properties. Note that this group
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 35

also provides a function for determining the specific-heat for air at constant pressure (cp)
which can be used instead of the function provided by the ideal-gas group.

Table 2.14. Properties of air at 1 atm pressure given the temperature in oC


# Function Output Output unit
1 Airrho_T Density (ρ) kg/m3
2 Aircp_T Specific heat (cp) J/Kg.oC
3 Airk_T Thermal conductivity (k) W/m.oC
4 Airdf_T Thermal diffusivity (α) m2/s
5 Airdv_T Dynamic viscosity (μ) kg/m·s
6 Airkv_T Kinematic viscosity (ν) m2/s
7 AirPr_T Prandtl number (Pr) -

2.3. Using Thermax property functions in Excel formulae


Excel comes with a number of add-ins that extend the application’s functionality such as
the Solver add-in [11]. Active add-ins are automatically loaded when Excel starts up, but
before Thermax can be recognised by Excel you have to install it in your computer. To
do that, open the Thermax1 file and then save it in your computer as an “Excel Add-
in”. Recent Excel versions locate all add-ins in a certain folder in the computer that
depends on the version of Excel you are using. Excel automatically directs you to the
appropriate location. Save the add-in and restart Excel in order to activate it. Open a new
Excel sheet and then do the following:

1. Go to File and then click Options.


2. Select Add-Ins. From the Manage ribbon at the bottom of the menu select Excel
Add-ins and then press Go.
3. The pull-down menu shown in Figure 2.2 will appear to you. To add Thermax
to the add-ins menu, tick (√) the corresponding box. If for any reason you saved
the add-in in a location that is different from the default folder, then click on
Browse and search for it in the destination folder and select it.

Having installed the add-in, its functions can be used in Excel's formulae just like its
built-in functions. The following sections illustrate two methods for using the add-in
functions in Excel formulae.

2.3.1. Accessing Thermax functions via the Function Wizard


This method suits the first-time users. To illustrate the method, let us start a formula by
entering the equal sign (=) in any cell (say, cell B2). If you now press the fx button in the
formula ribbon, the Function Wizard shown in Figure 2.3 will come up. The Function
Wizard firstly lists the various categories of built-in functions provided by Excel, e.g.,
financial, mathematical, statistical, etc. Scroll down the list of function categories and
select the User-defined functions. Then, all the functions provided by Thermax will be
listed alphabetically as shown in Figure 2.4.
36 Mohamed M. El-Awad

Figure 2.2. Adding Thermax to the menu of Excel add-ins

Figure 2.3. Finding the add-in user-defined functions in the Function Wizard
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 37

Figure 2.4. Thermax functions listed alphabetically in the User Defined category

The first function in the list of Figure 2.4, Air_Data, is an auxiliary function that stores
the data for the seven thermophysical properties of air at standard atmospheric pressure.
This function is called by other functions in the Air-group in order to determine the values
of their respective properties at the required temperature. To start using the add-in
functions, let us use it to determine the thermal conductivity (k) of air at 25oC. To do that,
go down the list and select the function named Airk_T. Upon pressing the OK button,
the Function Arguments box shown in Figure 2.5 will appear to you.

Figure 2.5. The Function Arguments box for the “Airk_T” function

Figure 2.5 shows that this function has one input parameter, which is the temperature in
o
C “TempC”, and gives a brief description of its intended use. Fill the form by entering
38 Mohamed M. El-Awad

the given value of the temperature, 25, as shown in Figure 2.6. Note that the formula
ribbon now shows the formula in cell B2, which is “=Airk_T(25)”. The form also shows
the calculated value of k, which is 0.02551 W/m.oC. When you press the “OK” button,
this value will appear in the cell B2. Check this value with the tabulated data.

Figure 2.6. Using the function “Airk_T” to determine the thermal conductivity of
atmospheric air at 25oC

2.3.2. Direct use of Thermax functions in Excel formulae


It is not necessary to follow the lengthy procedure described above by using the Function
Wizard in order to use the add-in property functions. After some practice one can simply
type in the formula that contains the property function in the required Excel cell. As
mentioned earlier, the style adopted for naming Thermax functions makes it easy to
identify the required function and the required input without having to memorise these
for all the functions. For illustration, suppose that we want to determine the temperature
of carbon-dioxide (CO2) at which the value of its enthalpy (h) equals 750 kJ/kg. In this
case, we need to use the function in the “Gas” group that determines the temperature for
a given enthalpy value. Therefore, we start an Excel formula by typing the equal sign “=”
in any cell and then type the prefix “Gas”. As shown in Figure 2.7, as soon as we type
the prefix "Gas" after the equal sign, the user-interface will display all Thermax functions
in the ideal-gases group for us to select from.

Since the property we want to find is the temperature, which the Gas-group functions
require in absolute degree, we continue the name of the function by adding the letters
“TK” immediately after the three-letter prefix "Gas" followed by an underscore. As
shown in Figure 2.8, the user-interface then lists only the five functions in Table 2.2 that
determine the gas temperature given h, Pr, s0, u, or vr. Since we want to find the
temperature from a known value of enthalpy, we have to select the “GasTK_h” function
as shown in Figure 2.9.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 39

Figure 2.7. Excel UI showing all the functions in the Gas-group

Figure 2.8. UI showing only the five functions that determine the temperature of an
ideal gas from known values of h, pr, s0, u, or vr

Figure 2.9. The required input for the GasTK_h function

Excel will then ask you to enter the required input parameters. This function requires the
name of the gas, which is “CO2”, and the value of enthalpy, which is 750 kJ/kg, as shown
40 Mohamed M. El-Awad

in Figure 2.10. If you are not sure about the required input parameters you can call the
Function Wizard by pressing the fx button in the formula ribbon and the form shown in
Figure 2.11 will appear to you. The form shows the two required input parameters for
you to provide. Pressing the “Enter” key or the “OK” button after entering the required
data, the function will calculate the corresponding temperature which is shown in Figure
2.12 as 817.5544K. This value can be checked by comparison with the tabulated data.

Figure 2.10. The required input for the GasTK_h function

Figure 2.11. The Function Wizard with the required input for the GasTK_h function

Figure 2.12. The temperature determined by the GasTK_h function


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 41

2.4. Analyses with variable specific heat using Thermax functions


The following example shows how Thermax functions in the Gas-group can be used for
applying the variable specific-heat method (the exact method) in thermodynamic
analyses.

Example 2.1. Thermodynamic analysis with the exact method


Figure 2.13 shows a well-insulated piston–cylinder device that initially contains 0.1 m3
of air at 100 kPa, 330K. Fifty kJ of heat is transferred to the air causing the air to expand
at constant pressure. Treating air as an ideal gas, determine the final temperature inside
the cylinder using: (a) the approximate constant-specific heat method, with cp= 1.005
kJ/kg.K, and (b) the exact variable specific-heat method with Thermax functions. Also,
determine the error of the approximate method in determining the final temperature.

Air
T1 = 330K
P1 = 100 kPa
V1 = 0.1 m3

Heat 50 kJ
Figure 2.13. Schematic for Example 2.1

(a) Solution with the approximate method


Using the approximate constant specific heat method, the final temperature is determined
from:

Q
T2  T1  (2.2)
m  cp

Where m is the mass of air in the piston-cylinder device, cp is the average value of the
specific heat at constant pressure, and Q is the amount of heat added. The mass of air is
first calculated from the ideal-gas law:

P1V1
m (2.3)
R  T1
42 Mohamed M. El-Awad

Where R is the gas constant for air, R = 0.287 kJ/kg.K. Substituting for P1, V1, R and T1
in Equation (2.3), leads to m = 0.105585 kg and substituting for m in Equation (2.2) gives
T2 = 801.194K.

(b) Solution with the exact method


To apply the exact method by using Thermax functions, we have to determine the final
enthalpy, h2. From the first law of thermodynamics applied to the closed system that
undergoes a constant-pressure process, h2 is given by:

h2  h1  Q / m (2.4)

Once h2 is found, the final temperature, T2, can be determined by using the function
GasTK_h in the Gas-group. Figure 2.14 shows the Excel sheet developed for this
example. The given data are inserted at the left-hand side of the sheet together with the
gas constant (R_) and specific heat (cp) for air.

Figure 2.14. Excel sheet developed for Example 2.1

The calculations part is divided into two parts that determine the final temperature
according to the approximate method and the exact method using the corresponding
equations given above. Figure 2.14 also reveals the formulae typed in these calculations.
Note that “cell-labelling” has been used in all the Excel sheets described in this book [12,
13]. As the figure shows, the answer found by the approximate method for T2 is 801.2K,
while the exact method determines the final temperature as 781.6K. Thus, the
approximate constant specific-heat method results in an error of 2.5%.

2.5. Closure
This chapter introduced the Thermax add-in and presented its seven groups of property
functions for ideal gases, saturated water and superheated steam, refrigerants, humid air
for psychrometric analyses, two aqua solutions for vapour-absorption refrigeration,
reacting substances, and air at atmospheric pressure. Thermax property functions are
named in a way that helps the user to find the appropriate function via Excel’s user-
interface and function wizard. The chapter showed how the add-in functions can be used
in Excel formulae and illustrated the use of the ideal-gas functions for thermodynamic
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 43

analyses with the exact variable specific-heat method. There are a number of auxiliary
functions that are provided by the add-in groups but do not appear in the respective tables.
Some of these appear in the following chapters where their use is explained.

Thermax has been developed for educational purposes with the aim of making Excel
suitable for various types of thermofluid analyses. Therefore, for developing its property
functions the priority has been given to their variety rather accuracy. Almost all the data
and formulae used for the development of Thermax property functions have been
obtained from the open literature; the only exception being the formulae for superheated
refrigerants. In this respect, it should be emphasised that the functions for superheated
refrigerants only apply for subcritical pressures. Appendix C verifies these functions for
R134a by comparing their estimations with the data given by ASHRAE [2] and El-Awad
et al [14] verified them for two halocarbon refrigerants (R410A and R1234yf) and two
natural refrigerants (R290, and R744).

As shown in the following chapters, the automation of property calculations by using


property functions enables Excel to be used for iterative solutions and optimisation
analyses of various types of energy-utilisation systems using its two What-if analysis
tools; Goal-Seek and Solver [12, 13]. The scope of thermodynamic analyses with the
Excel-based platform can be widened by using a number of other educational and
research-oriented Excel add-ins that have been developed by academic institutions and
individuals [15,16].

References
[1] Y. A. Cengel and M. A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[2] American Society of Heating, Refrigeration and Air-Conditioning Engineers,
(ASHRAE), Handbook of fundamentals, Atlanta, 2017.
[3] T.F. Irvine Jr. and P.E. Liley. Steam and gas tables with computer equations.
Orlando, (FL): Academic Press, 1984.
[4] J. Patek and J. Klomfar. A computationally effective formulation of the
thermodynamic properties of LiBr–H2O solutions from 273 to 500 K over full
composition range. International Journal of Refrigeration, Vol. 29, 2006, pp.566–
578.
[5] M.R. Patterson and H. Perez-Blanco, “Numerical fits of the properties of lithium-
bromide water solutions”, ASHRAE Trans. Volume 94, Issue 2, pp. 2059-2077,
1988.
[6] J. Patek and J. Klomfar, "Simple functions for fast calculations of selected
thermodynamic properties of the ammonia-water system". International Journal of
Refrigeration, Vol. 18, No. 4, 1995, pp. 228 -234.
[7] D-W Sun. "Comparison of the performances of NH3-H2O, NH3-LiNO3 and NH3-
NaSCN absorption refrigeration systems". Energy Convers. Mgmt, Vol. 39, No. 5/6,
1998, pp. 357-368.
44 Mohamed M. El-Awad

[8] M.A.I. El-Shaarawi, S.A.M. Said, M.U. Siddiqui, "New correlation equations for
ammonia-water vapor liquid equilibrium (VLE) thermodynamic properties",
ASHRAE Transactions, 2013, pp. 293-298.
[9] P. Bourseau and R. Bugarel, “Absorption-diffusion machines: comparison of the
performances of NH3-H2O and NH3-NaSCN”. International Journal of
Refrigeration, Vol. 9, Issue 4, pp. 206-214.
[10] Y. A. Cengel and A. J. Ghajar. Heat and Mass Transfer: Fundamentals and
Applications. McGraw Hill, 2011.
[11] Frontline Systems, internet: http://www.Solver.com/ (Last accessed November 23,
2015).
[12] M.M. El-Awad, Computer-Aided Thermofluid Analyses using Excel -
https://www.academia.edu/40015718/Computer_Aided_Thermofluid_Analyses_
using_Excel
[13] M.M. El-Awad, Optimisation Analyses of Fluid-Thermal Systems using Excel -
https://www.academia.edu/36307413/Optimisation_Analyses_of_Fluid_Thermal
_Systems_using_Excel
[14] M.M. El-Awad, M.S. Al Nabhani, K.S. Al Hinai, A. Younis, Development and
Validation of an Excel Add-In for determining the Properties of Various
Refrigerants, Proceedings of the 1st National Conference on Applied Science,
Engineering & Technology 2019, CASET – 2K19, 11th June 2019, Ibri, Oman.
[15] D. G. Goodwin, "TPX: thermodynamic properties for Excel",
http://termodinamicaparaiq.blogspot.com/p/tpx_12.html (Last accessed July 11,
2019)
[16] The University of Virginia, Free Excel/VBA Spreadsheets for Thermodynamics:
Rankine, Brayton, Otto and Diesel Cycles, Internet:
http://www.faculty.virginia.edu/ribando/ modules/xls/Thermodynamics/ (Last
accessed July 1, 2019)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 45

Exercises
1. Complete the following table by specifying the usage of the given Thermax function
according to its name:

# Function Output property Input /properties


1 Gascp_TK(“CO2”,300)
2 WatTsat_P(200)
3 WatT_Ph(200, 3487.7)
4 RefPsat_T(“R134a”,0.0)
5 PsyTs_P(100)
6 Nh3Pr_TX(200,80)
7 Airdv_T(500)
8 Chm_f2(“HDL”,”QL”)

2. Name the appropriate Thermax functions that determine the following properties:
# Task Thermax function
o
1 The enthalpy of saturated liquid water at 25 C
2 The internal energy of oxygen at 300K
3 The entropy of superheated steam at 10 kPa, 200oC
4 The absolute temperature at which the relative pressure of
air is equal to 15
5 The saturation pressure for R-134a at -15oC
6 The dry-bulb temperature of humid air at P = 101 kPa, ϕ =
50%, ω = 0.01
7 The enthalpy of ammonia-water solution at 30oC, X= 50%
8 The higher heating value for gasoline

3. Using Thermax functions, determine the values of the following fluid properties:
# Fluid Given properties Required property Value
1 Saturated water 100 kPa, x = 0.5 Enthalpy (h)
2 Air 700K Relative volume (vr)
o
3 Steam 100 C, x = 1.0 Entropy (s)
4 CO2 450K Specific heat (cp)
5 Hydrogen h = 2000 kJ/kg Temperature (T)
46 Mohamed M. El-Awad
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 47

Chapter 3

Analyses of gas power cycles


48 Mohamed M. El-Awad

This chapter illustrates the use of Thermax functions for analysing two important gas
power cycles which are the Brayton cycle for gas turbines and the Otto cycle for spark-
ignition internal-combustion engines. The results obtained for the Brayton cycle by using
the add-in functions for ideal-gases (Gas) are compared to those obtained by using the
“Ideal-Gas” add-in developed at the University of Alabama [1]. For the Otto cycle,
energy as well as exergy analyses are presented. The values obtained for the key
parameters of the cycles are verified against the relevant values given in standard
thermodynamics textbooks.

3.1. The ideal Brayton cycle


The Brayton cycle is the ideal cycle for gas-turbines. Unlike real gas-turbine cycle, it is
a closed cycle with a constant mass of the working fluid going through the system’s
components as shown in Figure 3.1. The cycle consists of four processes; compression
in an ideal gas compressor, heat-addition from a high-temperature source, expansion in
an ideal gas turbine, and heat-rejection to a low-temperature sink. Figure 3.2 shows the
Brayton cycle on a T-s diagram.
qin

Heat
exchanger
2 3
wt

wc

Compressor Turbine

1
Heat 4
exchanger

qout

Figure 3.1. Schematic diagram of the closed-cycle gas-turbine

Although the working fluid can be any suitable gas, in the following analysis it is assumed
to be air. Referring to Figure 3.2, the compressor work (wc) and turbine work (wt), per
unit mass flow rate of the air, are determined from:

wc  h2  h1  (3.1)

wt  h3  h4  (3.2)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 49

T 3

1
s

Figure 3.2. T-s diagram for the ideal Brayton cycle

Where h1, h2, h3, and h4 are values of the enthalpy at states, 1, 2, 3, and 4, respectively.
The amount of heat addition (qin) per unit mass flow of the air is given by:

qin  h3  h2  (3.3)

Therefore, the net work (wnet), back-work ratio (BWR), and thermal efficiency (η) of the
cycle are determined from:

wnet  wt  wc (3.4)

BWR  wc / wt (3.5)

  wnet / q in (3.6)

Given the compressor’s inlet temperature, T1, and the turbine’s inlet temperature, T3, the
discharge temperature from the compressor (T2) and the discharge temperature from the
turbine (T4) are determined by using the variable specific-heat method of analysis from
the corresponding relative pressures Pr2 and Pr4 given by:

P2
Pr 2  Pr1  (3.7)
P1
P
Pr 4  Pr 3  4 (3.8)
P3
50 Mohamed M. El-Awad

Where Pr1 and Pr2 are the relative pressures at states 1 and 2, respectively. For the ideal
cycle without pressure losses, P3 = P2 and P4 = P1. The following example illustrates the
procedure for using Thermax property functions for analysing the cycle with Excel. The
example is based on Example 9.5 in Cengel and Boles [2].

Example 3.1. Analysis of the simple ideal Brayton cycle


A gas-turbine power plant that operates on an ideal Brayton cycle with air as the working
fluid has a pressure ratio of 8. The gas temperature at the compressor inlet is 300 K and
at the turbine inlet is 1300 K. Determine (a) the net work per unit mass flow rate of the
working fluid, (b) the back work ratio, and (c) the thermal efficiency.

Solution
Figure 3.3 shows the Excel sheet developed for this example using Thermax functions.
The figure shows the general layout adopted in this book for sheet arrangement according
to which the sheet is divided from left to right into three main parts: (i) Input data, (ii)
Intermediate calculations, and (iii) Final results. In the present case, the input data include
the given values of the two temperature values at states 1 and 3 (T_1 and T_3) and the
pressure ratio (rp). The middle part of the sheet, which is reserved for the calculation of
intermediate results, shows the calculated values of the two unknown temperatures T_2
and T_4 and the enthalpy values at the four states in the cycle. The final results are the
compressor work (w_c), the turbine work (w_t), the heat input (Q_in), the back work
ratio (BWR), and thermal efficiency (η) as shown on the right side of the sheet.

Figure 3.3. The Excel sheet developed for Example 3.1

Figure 3.4 shows the formula entered in each cell of the calculations part using Thermax
functions. Note that “cell-labelling” has been used [3]. Only three Thermax functions
from the Gas-group have been used in the Excel model with air as the gas which are:
GasPr_TK, GasTK_Pr, and Gash_TK. For the purpose of comparison, a similar Excel
sheet was developed by using the IdealGas add-in developed at the University of
Alabama [1]. Figure 3.5 shows the Excel sheet and Figure 3.6 reveals the function used
in it. Note that the IdealGas functions have the advantage of allowing two systems of
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 51

units to be used, but they can only be used for air. By comparison, Thermax functions
only use the “SI” system but they can be used for different gases.

Figure 3.4. Thermax functions used in the Excel sheet of Example 3.1

Figure 3.5. Excel sheet for Example 3.1 using the IdealGas add-in

Figure 3.6. IdealGas functions used in the Excel sheet of Example 3.1

Table 3.1 shows the values determined by the two property add-ins for the key cycle
parameters and their corresponding values given by Cengel and Boles [2]. The figure of
52 Mohamed M. El-Awad

the table shows only minor differences between the values obtained by the property add-
ins and those given by Cengel and Boles [2].

Table 3.1. Verification of the two add-ins’ results with those of Cengel and Boles [2]
Parameter Cengel and Boles [2] Thermax IdealGas
wc 244.16 243.69 243.69
wt 606.60 606.01 606.01
qin 851.62 852.52 852.52
BWR 0.403 0.402 0.402
 0.426 0.425 0.425

3.2. The regenerative Brayton Cycle


Gas turbines can sustain higher combustion temperatures than those typically met in
internal-combustion engines, but the high combustion temperatures lead to high exhaust
temperatures as well. By adding a regenerator to the simple gas-turbine system, as shown
in Figure 3.7, the energy in the hot exhaust gases that can reach more than 500 oC is
utilised to preheat the compressed air before it goes to the combustion chamber; a process
called regeneration. Regeneration reduces the fuel combustion and improves the plant’s
thermal efficiency, but the cost of the heat-exchanger increases that of the modified
system. The economic benefit due to regeneration depends not only on the cost of the
heat-exchanger, but also on the system’s pressure ratio. While the cost of the heat-
exchanger depends on its size and effectiveness, there is a certain value for the pressure
ratio that maximises the system’s thermal efficiency.

Regenerator
6

C.C. 4
5 3
2
1
G
Compressor Turbine

Figure 3.7. Schematic diagram of the regenerative gas-turbine

Figure 3.8 shows the T-s diagram for the regenerative Brayton cycle assuming zero
pressure losses in the combustion chamber and heat-exchanger. Due to friction losses in
both compressor and turbine, the actual exit temperatures from these devices (i.e., T2 and
T4) are different from the ideal values obtainable by the isentropic compression and
expansion processes (T2s and T4s).
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 53

T 3

5
2
2s
4
4s

1
s
Figure 3.8. T-s diagram for the regenerative Brayton cycle

As for the simple Brayton cycle considered earlier, the thermal efficiency (η) is given by:

  wnet / q in (3.9)

Where, wnet is the net output work from the plant and qin is its heat input. The net output
work for the gas-turbine power plant is the difference between the turbine work output
(wt) and the compressor work input (wc). Using the numbering scheme of Figures 3.7 and
3.8 to indicate the state of the working fluid entering and leaving the compressor and
turbine, the net output work in terms of fluid enthalpies is given by:

wnet  h3  h4   h2  h1  (3.10)

The heat input to the power plant is given by the difference in enthalpies of the
combustion gases and the preheated air entering the combustion chamber, i.e.:

qin  h3  h5  (3.11)

Given the temperature and pressure of air at the compressor inlet, the ideal relative
pressure after the compressor (Pr,2s) is determined from [2]:

P2
Pr 2 s  Pr1  Pr1 .r p (3.12)
P1

Where, rp stands for the pressure ratio. Pr2s is then used to find T2s, which is the
temperature after an ideal isentropic compression process. Once T2s is found, h2s can be
determined by using the property function. The actual enthalpy (h2) is then found from:
54 Mohamed M. El-Awad

h2  h1  h2 s  h1  /  c (3.13)

Where, ηc is the compressor’s isentropic efficiency. Similarly, the relative pressure after
the expansion process (Pr,4s) is determined from:

P4
Pr 4 s  Pr 3  Pr 3 / r p (3.14)
P3

Pr4s is then used to find the temperature after an ideal expansion process T4s, from which
h4s can be determined. The actual enthalpy (h4) is then found from:

h4  h3  h3  h4 s    t (3.15)

Where, ηt is the turbine’s isentropic efficiency. The temperature of the compressed air
entering the combustion chamber (T5) and, therefore, the saving in fuel consumption,
depends on the effectiveness of the regenerator (ε) which is defined as:

  T5  T2  / T4  T2  (3.16)

Given an estimate for ε, Equation (3.16) can be rearranged to get:

T5  T2   (T4  T2 ) (3.17)

Example 3.2. Analysis of the regenerative Brayton cycle


A gas-turbine power plant operates on a regenerative Brayton cycle with air entering the
compressor at 100 kPa, 300 K. Combustion gases leave the combustion chamber at 1400
K. The regenerator effectiveness is 80% and the isentropic efficiency of both the
compressor and the turbine is 80%. Determine the pressure ratio that yields the maximum
net work output and the maximum thermal efficiency of the plant.

Solution
Figure 3.9 shows the Excel sheet prepared for this example using Thermax functions.
The input data shown in the left-side of the sheet includes the air intake temperature (T_1)
and pressure (P_1), the pressure ratio (Pratio), and the combustion temperature (T_3).
The data part also includes the isentropic efficiency of the compressor (Eff_c), the
isentropic efficiency of the turbine (Eff_t), and the regenerator effectiveness (Eff_regen).
The middle part of the sheet shows the calculated values of T2 and T4. The figure shows
the enthalpy values as obtained by using Thermax functions. The formula used in each
cell is inserted next to it. The right-side of the sheet shows the main results, which include
values of the input compression work, the output expansion work, the net work output,
the heat input, and the thermal efficiency.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 55

Figure 3.9. The Excel sheet developed for the regenerative Brayton cycle

The sheet shows the results at a pressure ratio of 10 at which the corresponding values of
the net output work and thermal efficiency are 216.19 kJ/kg and 0.346, respectively. The
value of the pressure-ratio (Pratio) in cell C4 was changed from 2 to 16 and Figure 3.10
shows the calculated values for the net specific work and thermal efficiency plotted
against the pressure ratio. The figure shows that the net output work and thermal
efficiency reach their maximum values at different pressure ratios. While the maximum
net work output occurs at a pressure ratio of about 8, the maximum thermal efficiency
occurs much earlier at a pressure ratio of about 5 or less. Note that exact values of the
intermediate pressure that maximise the net wok or thermal efficiency of the cycle can
be determined by using the Solver add-in (refer to Exercise 1).

250 0.45
0.4
Net specific work (kJ/kg)

200 0.35
Thermal efficiency

0.3
150
0.25
0.2
100
w_net 0.15
50 0.1
η
0.05
0 0
0 5 10 15 20
Pressure ratio
Figure 3.10. Variation of the net specific work and thermal efficiency of the
regenerative Brayton cycle with the turbine’s pressure ratio
56 Mohamed M. El-Awad

3.3. Energy analysis of the Otto cycle


The Otto cycle and the Diesel cycle are two ideal cycles used for modelling the cycles in
internal combustion (I.C.) engines. The Otto cycle is used for modelling spark-ignition
I.C. engines, while the Diesel cycle is used for modelling compression-ignition I.C.
engines. Both cycles adopt “air-standard” assumptions according to which the working
fluid is considered to be air which is treated as an ideal gas. Figure 3.11 shows the four
processes that constitute the Otto cycle:

Process 1-2: Adiabatic and reversible compression of air


Process 2-3: Constant-volume heat addition
Process 3-4: Adiabatic and reversible expansion of air
Process 4-1: Constant-volume heat rejection

The P-v diagram of the Otto cycle is shown in Figure 3.12. Processes 1-2 and 3-4 are
isentropic processes because they are both assumed to be adiabatic and reversible.
qin qout

Air 2 Air Air


2,3 3
Air

1 4 4,1

1-2: Isentropic 2-3: Heat addition 3-4: Isentropic 4-1: Heat rejection
compression at constant v expansion at constant v

Figure 3.11. The Otto cycle (Adapted from Cengel and Boles [2])

P 3

qin

4
2
qout

TDC BDC v

Figure 3.12. P-v diagram of the Otto cycle (Adapted from Cengel and Boles [2])
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 57

The analytical model


The analysis of the Otto cycle that aim to determine the thermal efficiency of the cycle is
called an “energy analysis”. For the ideal Otto-cycle shown in Figure 3.12, we have [4]:

v1  RT1 / P1 (3.18)

v 2  v1 / r (3.19)

v3  v 2 (3.20)

v4  v1 (3.21)

Where, R is the gas constant (for air R= 0.287 kJ/kg.K) and r is the compression ratio of
the cycle (for modern spark-ignition engines 8 ≤ r ≤ 11). Applying the variable specific
heat method for or the isentropic compression and expansion processes:

vr 2 v2
  1/ r (3.22)
v r1 v1
vr 4 v4
 r (3.23)
v r 3 v3

Where, vr is the relative specific volume. The last two relationships can be used to
determine the two temperatures after the compression and expansion processes. The
amount of heat added in process 2-3 per kg of the working fluid (qin) is calculated from:

qin  u3  u 2  (3.24)

The amount of heat removed per kg of the working fluid (qin) is calculated from:

qout  u 4  u1  (3.25)

Applying the first-law of thermodynamics, the net work from the cycle is given by:

wnet  qin  qout (3.26)

Therefore, the thermal efficiency of the Ott cycle (ηOtto) is given by:

wnet
 Otto  (3.27)
qin
58 Mohamed M. El-Awad

The cycle’s mean-effective pressure (MEP) is defined as:

wnet wnet
MEP   (3.28)
v max  v min v1  v 2

Although the above analytical model is based on the usual air-standard assumptions, by
adopting the variable specific-heat method, Equations (3.22) and (3.23), it yields more
realistic results than those obtainable by adopting the approximate constant specific heat
method (refer to Example 2.1). The following example shows how the property functions
provided by Thermax and IdealGas [1] can be used to be used for its application.

Example 3.3. Energy analysis of the Otto cycle


An ideal Otto cycle has a compression ratio of 8. At the beginning of the compression
process, air is at 100 kPa and 17°C. During the constant-volume heat-addition process,
800 kJ/kg of heat is transferred to air. Accounting for the variation of specific heats of air
with temperature, conduct an energy analysis of the cycle to determine:

(i) the net work output and thermal efficiency,


(ii) the mean effective pressure for the cycle.

This example is based on Examples 9-2 in Cengel and Boles [2].

Solution
Figures 3.13 and 3.14 show the Excel sheets developed for analysing the Otto cycle by
using property functions for ideal gases provided by Thermax and IdealGAs,
respectively. The sheets show the formulae used in the calculation parts and the formula
window in each sheet reveals the formula used for calculating the mean effective pressure
according to Equation (3.28). Since the IdealGas add-in does not provide a function for
determining the gas temperature from its internal energy, the temperature T_3 in Figure
3.14 has been found by using the Goal-Seek command of Excel. Table 3.2 compares the
values obtained by the two add-ins to those given by Cengel and Boles [2]. The figures
in the table show good agreement between the three solutions.

Figure 3.13. Excel sheet developed for Example 3.3


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 59

Figure 3.14. Excel sheet developed for Example 3.3 using the IdealGas add-in

Table 3.2. Key parameters in the energy analysis


Parameter Cengel and Boles [2] Thermax IdealGas
T2 652.4 648.7 648.7
T3 1575.1 1572.6 1572.6
T4 795.6 793.1 793.1
P2 1799.7 1789.6 1789.6
P3 4345.0 4338.2 4338.2
qout 381.83 381.91 381.91
wnet 418.17 418.09 418.09
ηOtto 0.523 0.523 0.523
MEP 574.0 574.1 574.1

Using the approximate method of analysis with fixed values of the specific heat, it can
be shown that the thermal efficiency of the Otto cycle is given by [4]:

 otto  1  1 / rc k 1 (3.29)

Where, k is the ratio of specific heats cp/cv for air. Substituting rc = 8 and k = 1.4 in
Equation (3.29), the calculated thermal efficiency is 0.565. It should be noted that values
of the thermal efficiency obtained by the exact and the approximate methods are both
exaggerated compared to that of actual engines, but that of the variable specific heat
method, which is 0.523, is less inaccurate.

3.4. Exergy analysis of the Otto cycle


Energy analyses, such as those presented in the previous sections, evaluate the general
performance of energy-utilisation systems using overall performance indicators such as
the thermal efficiency for power-producing systems and the coefficient of performance
(COP) for refrigeration systems. By comparison, exergy analyses which are based on the
second-law of thermodynamics enable the locations, types, and true magnitudes of waste
and loss to be determined. Therefore, exergy analyses can be used in design analyses of
thermofluid systems to further the goal of achieving more efficient use of resources [5].
60 Mohamed M. El-Awad

The analytical model


Exergy (ϕ) is a thermodynamic property that measures the ability of the working fluid to
do useful work. Per unit mass in a closed system like that of the engine shown in Figure
3.11, exergy is given by [2]:

V2
  u  u 0   T0 s  s 0   Po v  v0    gz (3.30)
2

Where u is the internal energy and u0, s0, and P0, respectively, refer to the values of the
internal energy, entropy and pressure at the dead state which is the surroundings.
Neglecting changes in the kinetic and potential energies, and taking into consideration
that v3 = v2 in this case, the exergy input in process 2-3 is determined from:

input  3  2   u3  u2   T0 s3  s2  (3.31)

Exergy of the working fluid can be destroyed in a process because of irreversibilities such
as friction losses and heat-transfer over a finite-temperature difference. In general, the
exergy destruction (xdest) in a process is determined from:

xdest  T0 s gen  T0 ssys  sin  sout 


 q q 
 T0  s 2  s1 sys  in  out 

(3.32)
 Tb,in Tb,out 

Where, sgen refers to the entropy generated in the process. In the Otto cycle, the processes
1-2 and 3-4 are both adiabatic and reversible. Therefore, for these two processes:

x dest  0 (3.33)

However, the processes 2-3 and 4-1 both involve heat transfer with the surroundings and
that makes them externally irreversible processes. While process 2-3 involves heat
addition only, process 4-1 involves heat rejection only. For these two processes, exergy
destructions are obtained from:

 q 
x dest  T0  s3  s 2   in  Process 2-3 (3.34)
 TH 

 
xdest  T0  s1  s4   out
q
 Process 4-1 (3.35)
 TL 
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 61

The exergy destruction of the whole Otto cycle is the sum of the exergy destructions in
the heat-addition and heat-rejection processes. Since s 2  s1 and s 4  s3 , the exergy
destruction in the cycle becomes:

q q 
xdest  T0  out  in  (3.36)
 TL TH 

This following example, which is based on Example 9-10 in Cengel and Boles [2],
verifies the relevant Thermax functions by comparing their calculations for the exergy
analysis with those given by Cengel and Boles [2].

Example 3.4. Exergy analyses of the Otto cycle


Accounting for the variation of specific heats of air with temperature, conduct an exergy
analysis for the Otto cycle considered in Example 3.3 to determine:

(i) the exergy destruction associated with the Otto cycle (all four processes as well
as the cycle), assuming that heat is transferred to the working fluid from a
source at 1700 K and heat is rejected to the surroundings at 290 K, and
(ii) the exergy of the exhaust gases when they are purged.

Solution
The two Excel sheets developed for Example 3.3 were extended for this exergy analysis
and Figure 3.15 shows the extension that is needed in the lower part of the sheet using
Thermax functions. Table 3.3 shows the functions used in this extension. The sheet using
the IdealGas add-in was extended in a similar way. The additional data needed for exergy
analysis include the temperature and pressure of the surroundings, temperature of the heat
source, and the sink temperature. By obtaining the required entropy values at the different
states from the energy part, the two sheets determine the exergy destructions in the four
processes of the cycle, the total exergy destruction, and the exergy lost in the heat-
rejection process. Table 3.4 compares the calculated values to those given by Cengel and
Boles [2]. The figures in the table show a good agreement between the results of the two
add-ins and those given by Cengel and Boles [2].

Figure 3.15. Excel sheet developed for Example 3.4 using Thermax functions
62 Mohamed M. El-Awad

Table 3.3. Thermax functions used in exergy analysis of the Otto cycle
Cell Cell name Formula
F30 Dels23 =Gass0_TK("air",T_2)-Gass0_TK("air",T_3)-
R_air*LN(P_2/P_3)
F31 Ed_23 =T_0*(-Dels23-Q_in/T_source)
F33 Dels14 =Dels23
F34 Ed_14 =T_0*(Dels14+q_out/T_sink)
I27 Ed_cycle =Ed_23+Ed_41
I29 v_0 =v_1
I30 u_0 =Gasu_TK("air",T_0)
I31 Dels40 =-Dels14
I33 E_4 =(u_4-u_0)-T_0*Dels40-P_0*(v_4-v_0)

Table 3.4. Results of the Otto cycle exergy analysis


Cengel and Boles [2] Thermax IdealGas
xdest ,23 82.2 82.8 82.79
x dest , 4 1 163.2 162.7 162.65
x dest ,Otto 245.4 245.4 245.44

4 163.2 162.6 162.65

The figures in Table 3.4 show that two thirds of the total exergy supplied to the engine is
lost during the heat-rejection process. Therefore, any design efforts that aim to minimise
the loss of exergy in the heat-rejection process can effectively improve the performance
of the Otto cycle. For example, the lost exergy can be used in a cogeneration system that
utilises the heat for producing steam for industrial applications or for air-conditioning
purposes. The figures in Table 3.4 also show that the exergy destruction in the heat-
addition process are about one third of the total exergy destruction. While increasing the
heat-addition temperature will improve the thermal efficiency of the cycle, it will increase
the rate of exergy destruction in this process. Therefore, this temperature requires a
careful consideration of the two factors among other practical considerations.

3.5. Closure
This chapter illustrated the use of Thermax property functions for the analyses of two
basic power-generation cycles, which are the Brayton cycle and the Otto cycle. With
respect to such gas cycles, property functions enable the exact variable specific-heat
method to be used instead of the approximate constant-specific heat method. The
analyses show that the functions provided by Thermax give identical results to those
given by the IdealGas add-in. Thermax provides additional functions not provided by
IdealGas such as the GasTK_u and GasTK_h functions that determine the temperature
from a given value of internal energy or enthalpy, respectively.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 63

References
[1] The University of Alabama, Mechanical Engineering, Excel for Mechanical
Engineering project, Internet: http://www.me.ua.edu/excelinme/index.htm (Last
accessed July 19, 2018)
[2] Y. A. Cengel and M. A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 8th Edition, 2015
[3] M.M. El-Awad, Computer-Aided Thermofluid Analyses using Excel -
https://www.academia.edu/40015718/Computer_Aided_Thermofluid_Analyses_
using_Excel
[4] W.W. Pulkrabek, Engineering Fundamentals of the Internal Combustion Engine,
Second Edition, Pearson Prentice Hall, 2004.
[5] A. Bejan, G. Tsatsaronis, M. Moran, Thermal Design & Optimization, John Wiley
$ Sons, 1996.
[6] Frontline Systems, internet: http://www.Solver.com/ (Last accessed November 23,
2015).
[7] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley, & Sons. Inc. 2006

Exercises
1. Figure 3.10 shows that the net output work and thermal efficiency of the regenerative
Brayton cycle reach their maximum values at different pressure ratios; which is about
8 for the maximum net work output and about 5 or less for the maximum thermal
efficiency. Use Solver [6] to determine the exact values of the pressure ratios.

2. A regenerative gas turbine with intercooling and reheat operates at steady state. Air
enters the compressor at 100 kPa, 300 K with a mass flow rate of 5.8 kg/s. The
pressure ratio across the two-stage compressor is 10. The pressure ratio across the
two-stage turbine is also 10. The intercooler and reheater each operate at 300 kPa. At
the inlets to the turbine stages, the temperature is 1400 K. The temperature at the inlet
to the second compressor stage is 300 K. The isentropic efficiency of each
compressor and turbine stage is 80%. The regenerator effectiveness is 80%. Figure
3.P2 shows the T-s diagram of the regenerative gas turbine cycle.

Develop an Excel sheet to determine: (a) the thermal efficiency, (b) the back work
ratio, (c) the net power developed, in kW. Compare your solution with that of
Example 9.11 in Moran and Shapiro [7].

3. The air-standard Diesel cycle shown in Figure 3.P3 operates with a compression ratio
of 18. At the beginning of the compression process the temperature is 300 K and the
pressure is 0.1 MPa. The cutoff ratio for the cycle is 2. Using Thermax or any other
property add-in, develop and Excel sheet to determine (a) the temperature and
pressure at the end of each process of the cycle, (b) the thermal efficiency, (c) the
mean effective pressure, in MPa. Compare your solution with that of Example 9.2 in
Moran and Shapiro [7].
64 Mohamed M. El-Awad

T 6
8
5

7s 7
4s 4
2s 2 9
9s
3 10
1

s
Figure 3.P2. T-s diagram for the regenerative gas turbine with intercooling

P qin
2 3

4
qout

TDC Cut-off BDC v

Figure 3.P3. P-v diagram of the Diesel cycle


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 65

Chapter 4

Analyses of the Rankine cycle and the


combined Brayton-Rankine cycle
66 Mohamed M. El-Awad

Thermax property functions can be used with Excel to analyse various power-generation
cycles that use steam as the working fluid. This chapter illustrates the use of these
functions for analysing the simple Rankine cycle, Rankine cycle with superheat and
reheat, the regenerative Rankine cycle with one and two feed-water heaters, and the
combined Brayton-Rankine cycle. The results obtained by Thermax functions for the key
parameters of the cycles are verified against the relevant values given by Moran and
Shapiro [1] and Cengel and Boles [2].

4.1. The ideal simple Rankine cycle


The Rankine cycle is the ideal cycle for steam-turbine power plants. The basic plant
shown in Figure 4.1 consists of four components: a boiler, a turbine, a condenser, and a
pump. The working fluid, which is water, is pumped at a low temperature to the boiler
where it is heated to become saturated steam. The high-pressure steam is then taken to
the turbine where it expands to produce work. The steam is completely condensed in the
condenser before being pumped back to the boiler. Cooling water removes the heat
rejected in the condenser. Figure 4.2 shows the T-s diagram for the ideal cycle.

3
Turbine

Boiler 4

Condenser
2 Cooling
1 water
Pump

Figure 4.1. Schematic diagram of a simple steam-turbine power plant

T
3

1 4

s
Figure 4.2. T-s diagram for the ideal simple Rankine cycle
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 67

The pump-work (wp), turbine-work (wt), rate of heat addition in the boiler (qin), and rate
of heat rejection in the condenser (qout) per unit mass flow of the working fluid are given
by:

w p  h2  h1  (4.1)

wt  h3  h4  (4.2)

qin  h3  h2  (4.3)

qout  h4  h1  (4.4)

Where h1, h2, h3, and h4 are enthalpy values of the working fluid at the four states
indicated in Figures 4.1 and 4.2. The two enthalpy values h1 and h3 can be obtained
directly by using the relevant Thermax functions, but h2 and h4 have to be calculated
from:

h2  h1  v1 P2  P1  (4.5)

h4  h1  x4 h fg , P1 (4.6)

Where v1 is the specific volume of saturated liquid water at the condenser pressure, x4 is
the quality at point 4, and hfg,P1 is the enthalpy of vaporisation (the difference between
the specific enthalpy of saturated steam and saturated liquid water) at the condenser
pressure P1. The quality x4 is determined from:

s 4  s1
x4  (4.7)
s fg , P1

Where s1 and s4 are the specific entropy values at points 1 and 4, respectively, and sfg,P1
is the difference between the specific entropy of saturated steam and saturated liquid
water at the condenser pressure P1. Note that s4 = s3= sg,P2.

Once enthalpy values at the states 1 to 4 are determined, and the values of wp, wt, qin, and
qout are calculated, the different cycle parameters can be calculated as follows:

(a) The thermal efficiency (η):

  wt  w p  / qin (4.8)


68 Mohamed M. El-Awad

(b) The back work ratio (BWR):

BWR  w p / wt (4.9)

(c) The mass flow rate of the steam (kg/h):

W  1000  3600
 (4.10)
m
wt  w p

Where W is the net power of the plant in MW.

(d) The rate of heat transfer in the boiler, ( Q boiler ), into the working, in MW:


boiler  mqin / 3600
Q  (4.11)

(e) The rate of heat transfer from the condenser ( Q condenser ), in MW:


condenser  mqout / 3600
Q  (4.12)

(f) The mass flow rate of the cooling water in kg/h:

Q out  1000  3600


 cw  (4.13)
hcw,o  hcw,i 
m

Where hcw,i and hcw,o are the enthalpy values of the cooling water entering and exiting the
condenser, which are approximated with those of saturated water at the relevant
temperatures. The following example, which is based on Example 8.1 in Moran and
Shapiro [1], shows how the cycle can be analysed with Thermax property functions.

Example 4.1. Analysis of the simple and ideal Rankine cycle


A steam-turbine power plant that operates on the ideal Rankine cycle has a net power
output of 100 MW. Saturated steam enters the turbine at 8.0 MPa and saturated water
exits the condenser at 8 kPa. Determine (a) the thermal efficiency of the cycle, (b) the
back work ratio, (c) the mass flow rate of the steam, in kg/h, (d) the rate of heat transfer,
Qin, into the working fluid as it passes through the boiler, in MW, (e) the rate of heat
transfer, Qout, from the condensing steam as it passes through the condenser, in MW, (f)
the mass flow rate of the condenser cooling water, in kg/h, if cooling water enters the
condenser at 15oC and exits at 35oC. Also, study the effect of increasing the condenser
pressure from 8 to 20 kPa on the thermal efficiency, mass-flow rate of the steam, and
mass-flow rate of the cooling water.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 69

Solution
The Excel sheet developed for this example is shown in Figure 4.3. The left side of the
sheet stores the given data which include the condenser pressure (P_1), the boiler pressure
(P_2), the power of the cycle (Power), the temperature of the cooling water entering the
condenser (T_cwi), and the temperature of the cooling water leaving the condenser
(T_cwo). The central part of the sheet performs the intermediate calculations that
determine the values of enthalpy at the four states in the cycle. Table 4.1 shows the
formulae used in these calculations with the various property functions provided by
Thermax. For example, the enthalpy at point 1 (h1), which equals the enthalpy of saturated
liquid water at the condenser pressure, is obtained by using the add-in function Wath_Px
with the two input parameters being the condenser pressure and the quality x =0.

Figure 4.3. The Excel sheet developed for Example 4.1

Table 4.1. Excel formulae with Thermax functions in the sheet of Example 4.1
Cell Parameter Excel formula
E2 h_1 =Wath_Px(P_1,0.0)
E4 h_2 =h_1+Watv_Px(P_1,0.0)*(P_2-P_1)
E6 h_3 =Wath_Px(P_2,1.0)
E8 s_3 =Wats_Px(P_2,1.0)
E10 x_4 =(s_3-Wats_Px(P_1,0.0))/(Wats_Px(P_1,1.0) - Wats_Px(P_1,0.0))
E12 h_4 =h_1+x_4*(Wath_Px(P_1,1.0) - Wath_Px(P_1,0.0))

Cells H2 to H10 calculate the values of the pump-work (w_p), the turbine-work (w_t),
the net work (w_net = w_t – w_p), the rate of heat addition in the boiler (q_in), and rate
of heat rejection in the condenser (q_out). The final results given in the right-side of the
sheet include the thermal efficiency (η), the back-work ratio (BWR), the rate of heat
addition in the boiler (Q_boil), the rate of heat rejection in the condenser (Q_cond), and
the mass flow rate of the cooling water (m_cw). The formula bar in Figure 4.3 reveals
the formula in the cell K12 that calculates the mass flow rate of the cooling water from
Equation (4.13). The thermal efficiency of the simple Rankine cycle is 37%. Table 4.2
70 Mohamed M. El-Awad

compares the values obtained by Thermax functions and those given by Moran and
Shapiro [1] for the cycle parameters in the question. The figures show close agreements
between the two results.

Table 4.2. Verification of Thermax results for Example 4.1


Moran and Deviation
Parameter Thermax
Shapiro [1] (%)
(a) η 0.371 0.3709 -0.189
(b) BWR 0.0084 0.0084 0
(c) m (kg/h) 3.77x105 3.77 x105 0.265
(d) Qboiler (MW) 269.77 269.546 -0.083
(e) Qcondenser (MW) 169.75 169.546 -0.120
(f) m cw (kg/h) 7.3x10 6
7.3 x10 6
0

The Excel sheet was used to analyse the effect of increasing condenser pressure on the
cycle’s efficiency, mass flow rate of the steam and mass flow rate of the cooling water.
The resulting percentage changes of the three parameters are shown in Figure 4.4. The
figure shows that the thermal efficiency decreases by nearly 6%, while the mass flow
rates of both the steam and the cooling water increase by more than 10%.

15
Thermal efficiency

10 Flow rate of steam


Percentage change (%)

Flow rate of cooling


5
water

0
5 10 15 20 25
-5 Condenser pressure (kPa)

-10

Figure 4.4. Effect of condenser pressure on the cycle’s efficiency, mass flow rate of the
steam, and mass flow rate of the cooling water

4.2. The Rankine cycle with superheat and reheat


The thermal efficiency of the simple Rankine cycle can be improved by increasing the
boiler pressure and/or reducing the condenser pressure. However, both of these changes
increase the wetness fraction of the steam in the last stages of the turbine and may cause
erosion of the turbine blades in these stages. This problem can be solved by superheating
and/or reheating the steam. Figure 4.5 shows the schematic diagram of a steam-turbine
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 71

power plant with super-heating and reheating. Compared to the simple Rankine cycle
considered in Example 4.1, this cycle allows lower condenser pressures without
exceeding the maximum allowable wetness fraction in the last stages of the low-pressure
turbine. Figure 4.6 shows the T-s diagram for the cycle.

3
HP LP
Turbine Turbine
4
6
Steam
Generator
5
Condenser
2
1
Pump

Figure 4.5. Schematic diagram of a steam-turbine power plant with superheating and
reheating

T
3

4
2

1 6
s
Figure 4.6. T-s diagram for the ideal reheat cycle

Since there are two stages of heat addition and two stages of expansion, the relevant
equations for the total turbine-work output (wout) and heat-addition (qin) become:

wout  h3  h4   h5  h6  (4.14)

qin  h3  h2   h5  h4  (4.15)

Therefore, the net work-output and thermal efficiency of the cycle are given by:
72 Mohamed M. El-Awad

wnet  wout  win (4.16)

  wnet / qin (4.17)


Where win is the pump work input given by:

win  h2  h1  (4.18)

 , in kg per hour, is obtained from the total net power of


The mass flow rate of steam, m

the plant, W in MW, as follows:

W  1000  3600
m  (4.19)
wnet

The following example that shows how the superheated and reheated Rankine cycle can
be analysed by using Thermax is based on Example 8.3 in Moran and Shapiro [1].

Example 4.2. Analysis of an ideal reheat Rankine cycle


A power plant working on an ideal Rankine cycle with superheat and reheat uses steam
as the working fluid to produce a net power output of 100 MW. Superheated steam enters
the first-stage turbine at 8.0 MPa, 480oC, where it expands to 0.7 MPa. It is then reheated
to 440oC before entering the second-stage turbine and expanding to the condenser
pressure of 0.008 MPa. Determine:

(a) the thermal efficiency of the cycle,


(b) the mass flow rate of steam, in kg/h,
(c) the rate of heat transfer from the condensing steam as it passes through the
condenser, in MW.

Solution
Figure 4.7 shows the Excel sheet developed for this example. Based on the given data
shown on the left side of the sheet, Thermax property functions determine the enthalpy
(h), entropy (s), or quality (x) at the six fluid states in the cycle. The enthalpy values at
the different states are then used to determine the work and heat interactions according
to Equations (4.14) and (4.15). Table 4.3 shows the formulae and Thermax property
functions used in the determination of enthalpy and entropy or quality at each state point
and the values calculated by Thermax functions for the state and overall cycle parameters.
It appears from the table figures that all the values calculated by the present model deviate
from those given by Moran and Shapiro [1] by less than 1%.

This example reveals three advantage of the Rankine cycle with superheating and
reheating over the simple cycle considered in Example 4.1 that operates between the same
upper and lower pressure levels of 8 MPa and 8 kPa, respectively. The first advantage is
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 73

that the thermal efficiency of this cycle is 40.5%, while that of the simple cycle is only
37%. The second advantage of the cycle with superheating and reheating is the lower
moisture content in the last stages of the low-pressure turbine; which reduces the
possibility of turbine’s blade erosion. This is indicated by the higher value of the quality
at state 6, x6 = 0.939, compared to the corresponding value in the simple cycle, x4 = 0.675.
The third important advantage is that the mass flow rate in the cycle with superheating
and reheating is lower than that of the simple cycle; which reduces the cost of the boiler
and other components in the power plant.

Figure 4.7. Excel sheet developed for the analysis of the reheat Rankine cycle

Table 4.3. Verification of Thermax results for Example 4.2


Moran & Deviation
Name Excel formula with Thermax functions Thermax
Shapiro [1] (%)
h1 =Wath_Tx(T_1,x_1) 173.88 173.3569 -0.301
h2 =h_1+Watv_Tx(T_1,x1)*(P_2-P_6) 181.94 181.4163 -0.288
h3 =Wath_pT(P_3,T_3) 3348.4 3355.235 0.2041
s3 =Wats_pT(P_3,T_3) 6.6586 6.7001 0.6233
x4 =(s_3-Wats_Tx(T_4,x_4))/Watsfg_T(T_4) * 0.9895 0.9985 0.9096
h4 =Wath_Tx(T_4,0.0)+x_4*Wathfg_T(T_4) * 2741.8 2759.743 0.6544
h5 =Wath_pT(P_5,T_5) 3353.3 3353.811 0.0152
s5 =Wats_pT(P_5,T_5) 7.757 7.765 0.1031
x6 =(s_6-Wats_Tx(T_6,0.0))/Watsfg_T(T_6) *
0.9382 0.939 0.0853
h6 =Wath_Tx(T_6,0.0)+x_6*Wathfg_T(T_6) * 2428.5 2429.766 0.0521
 =w_out/Q_in 0.403 0.403 0
*
Watsfg_T and Wathfg_T return the entropy change and enthalpy of vaporisation for saturated
water at the given temperature, respectively. These two functions are not listed in Tables 2.3.
74 Mohamed M. El-Awad

4.3. Regenerative Rankine cycle with a single open feed-water heater


Regeneration in a Rankine cycle refers to the process of preheating the feed-water by the
energy of steam that is extracted from the turbine at an intermediate pressure.
Regeneration reduces the mass flow rate through the turbine and the work output of the
turbine, but improves the thermal efficiency. Figure 4.8 shows a steam-turbine power
plant that has a single open feed-water heater (FWH). Heat exchange between the bled
steam and the condensate can also take place in a single closed FWH or in multiple open
and closed FWHs. Figure 4.9 shows the T-s diagram for the system shown in Figure 4.8.

Figure 4.8. Schematic diagram of the steam power plant with regeneration (adopted
from Cengel and Boles [2])

1-y

Figure 4.9. T-s diagram for the Rankine cycle with regeneration
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 75

The main components of the plant shown in Figure 4.8 are a boiler (or steam generator),
a steam turbine, an open feed-water heater, a condenser, two feed-water pumps, and a
connecting pipes system. The working-fluid, which is water, undergoes a complete cycle
whereby it leaves the condenser as saturated liquid at Pc and then pumped by the first
pump (Pump I) to the pressure in the feed-water heater (P6). After mixing with the
extracted steam in the open FWH, the feed-water is pumped by the second pump (Pump
II) to the boiler pressure (Pb). In the boiler, it is heated to become superheated steam at
T2 before delivered to the steam turbine. The superheated steam expands in the steam
turbine to an intermediate pressure (P6).

Per each kilogram of steam leaving the boiler at point 5, a fraction y is extracted and
mixed with the condensate in the open feed-water heater. The rest of steam (1-y)
completes its expansion in the low-pressure turbine to the condenser pressure (Pc = P7
=P1). In the condenser, the steam rejects its heat to a stream of cooling water and, thus,
returns to its initial state as saturated liquid water at Pc. The open-feed-water heater does
the job of regeneration as well as deaeration.

The analytical model


For steady operation, the heat input, turbine’s work (wt), and input work consumed by
the two pumps (wPI and wPII) are given by:

Qin  h5  h4 (4.20)

wt  1  h5  h6   1  y h6  h7  (4.21)

wPI  v1 1  y P2  P1  (4.22)

wPII  v3 P4  P3  (4.23)

The enthalpies of the fluid leaving the turbine (h7) and two pumps (h2 and h4) are
determined from the corresponding values following isentropic-expansion and isentropic
compression (i.e. h7s, h2s and h4s) as follows:

h7  h6  (h6  h7 s )   t (4.24)

h2  h1  (h1  h2 s ) /  c (4.25)

h4  h3  (h3  h4 s ) /  c (4.26)
76 Mohamed M. El-Awad

Where, ηt and ηp are, respectively, the isentropic efficiencies of the two turbines and two
pumps. The fraction of steam extracted for regeneration (y) is obtained from mass and
energy balance over the feed-water heater:

h3  h2
y (4.27)
h6  h2
The working fluid at points 2 and 3 is saturated liquid water and, therefore, the enthalpy
values at these two points are those of saturated water hf at the corresponding pressures.
Water leaves the two pumps as compressed liquid and values of its enthalpies at the two
discharge points 2 and 3 can be estimated from:

h2 = h1 + wPI (4.28)

h4 = h3 + wPII (4.29)

Since the steam at point 5 is superheated, the value of its enthalpy can be determined
from its pressure and temperature. The steam at point 6 is also superheated and its entropy
is equal to that at state 5. Therefore, the value of it enthalpy can be obtained from the
pressure and entropy using the appropriate Thermax function.

The state of the steam at the last stage of expansion, i.e. point 7, lies in the saturated-
mixture region. Therefore, determining its enthalpy needs the dryness fraction at this
point to be known. Since the entropy at state 7 also equals that at state 5, the dryness
fraction (x7) is given by:

s 5  s f @ Pc
x7  (4.30)
s fg @ Pc

Entropy values sf and sfg at the condenser pressure can be determined by using the
appropriate Thermax functions.

An important consideration in the design of a power-generation plant using the


regenerative Rankine cycle is the selection of the intermediate pressure at which steam is
to be extracted for regeneration. As will be shown later, the maximum values of the
thermal efficiency and net work output are obtained at a certain value of this pressure.
The following numerical example that illustrates the use of Thermax functions to analyse
the cycle is based on an example given by Cengel and Boles [2].

Example 4.3. Regenerative Rankine cycle with a single open feed-water heater
Consider a steam power plant operating on the ideal regenerative Rankine cycle with one
open feed-water heaters. Steam enters the turbine at 15 MPa and 600oC and is condensed
in the condenser at a pressure of 10 kPa. Some steam leaves the turbine at an intermediate
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 77

pressure (Pi) and enters the open feed-water heater. Study the effect of the regeneration
steam pressure on the net specific output work and thermal efficiency of the plant and
determine the most suitable extraction pressure.

Solution
Figure 4.10 shows the Excel sheet prepared for this example. The input data shown in the
left-side of the sheet includes the boiler pressure (P_b) and temperature (T_b), the
condenser pressure (P_c), the intermediate pressure (P_i), the isentropic efficiency of the
two turbines (ηt), and the isentropic efficiency of the two pumps (ηP). Note that both ηt
and ηp have been assigned a value of 1 in this ideal case. The middle part of the sheet
shows the calculated enthalpy values obtained by using Thermax functions. Figure 4.11
shows the formulae used in the sheet.

Figure 4.10. The Excel sheet for Example 4.3

Figure 4.11. The formulae used in the Excel sheet for Example 4.3

The right-side of the sheet shows the main results, which include the values of y and x7,
the two input work components, the two output work components, the heat input, and the
thermal efficiency. Table 4.4 compares the values obtained for the key solution
parameters at an intermediate pressure Pi of 1.2 MPa with the corresponding values given
by Cengel and Boles [2]. The figures in the table show a good agreement between the
present values and those of Cengel and Boles [2] since the maximum deviation between
the two results is less than 2%.
78 Mohamed M. El-Awad

Table 4.4. Verification of Thermax results for Example 4.3


Parameter Cengel and Boles [2] Thermax Deviation (%)
h1 191.83 191.817 -0.00678
h2 193.03 193.020 -0.00518
h3 798.65 798.360 -0.03631
h4 814.37 814.070 -0.03684
h5 3582.3 3579.440 -0.07984
h6 2859.5 2893.458 1.18755
h7 2114.9 2132.459 0.83025
y 0.2271 0.2242 -1.27697
x7 0.8037 0.8113 0.94563
 0.463 0.4555 -1.61987

To study the effect of the intermediate pressure (Pi) on the system’s net output work and
thermal efficiency, these were calculated for different values of this pressure. Figure 4.12
shows the variation of the net output work and thermal efficiency with Pi. Clearly, the
net work output decreases continuously as the extraction pressure is increased, but the
thermal efficiency reaches a maximum value at approximately 1100 kPa. Therefore, the
intermediate pressure has to be lower than this value. The exact value of Pi must be an
economical compromise between the revenue generated due to fuel saving and that due
to increased power.

0.456 1400
0.4555 η
Net specific work output (w_net)

1350
0.455 w_net
Thermal efficiency (η)

0.4545 1300
0.454
1250
0.4535
0.453 1200
0.4525
1150
0.452
0.4515 1100
0 1000 2000 3000 4000
Extraction pressure (kPa)
Figure 4.12. Variation of the cycle’s net output work and thermal efficiency with the
intermediate pressure
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 79

4.4. Reheat-regenerative Rankine cycle with two feed-water heaters


Figure 4.13 shows a schematic diagram of a steam power plant that operates on a reheat-
regenerative Rankine cycle. Compared to the system shown in Figure 4.8, the present
power plant adds a low-pressure steam turbine, a reheater, a closed feed-water heater, a
third feed-water pump, and a mixing chamber. The additional cost of this equipment
should be compensated for by the anticipated increase in both the power and thermal
efficiency of the plant. Figure 4.14 shows the T-s diagram for the cycle in the ideal
situation.

Figure 4.13. A regenerative Rankine cycle with open and closed feed-water heaters
(adopted from Cengel and Boles [2])

The analytical model


In the following model it is assumed that per each kilogram of steam entering and leaving
the boiler and high-pressure turbine, a fraction y is passed through the closed feed-water
heater. Out of the remaining (1-y) kg that passes through the low-pressure heaters, a
fraction z is extracted for regeneration in the open feed-water heater. Therefore, the
fraction that passes through the condenser is (1-y-z). The specific work output, work
input, and heat input for this power plant are given by:

wout  1  (h9  h10 )  (1  y )(h11  h12 )  (1  y  z )(h12  h13 ) (4.31)

win  (1  y  z )(h2  h1 )  (1  y)(h4  h3 )  y (h7  h6 ) (4.32)


80 Mohamed M. El-Awad

Figure 4.14. T-s diagram for the reheat-regenerative Rankine cycle (adopted from
Cengel and Boles [2])

Qin  1  (h9  h8 )  (1  y )(h11  h10 ) (4.33)

Steam at points 9 and 11 is superheated and, therefore, enthalpy values at these two points
(h9 and h11) can be determined from the known values of pressure and temperature. Steam
at points 10 and 12 are also superheated at the reheat pressure and extraction pressure,
respectively. The other known property at these two points is entropy, which leads to
s10=s9 and s12=s11 in the case of ideal isentropic expansion. At point 13 the working fluid
is saturated liquid-vapour mixture at the condenser pressure. The entropy of the fluid
leaving the low-pressure turbine at point 13 is determined from the known condenser
pressure and entropy at point 13 which, for isentropic expansion, equals that at point 12,
i.e. s13=s12.

Since the water at points 1, 3 and 6 is saturated liquid at the condenser pressure, its
enthalpy at these points equal the enthalpies of saturated liquid water at the corresponding
pressures, i.e. h1=hf@P1, h3=hf@P3, and h6=hf@P6. Water leaves the three pumps as
compressed-liquid and its enthalpies at points 2, 4, and 7 are obtained from:

h2  h1  v1 ( P2  P1 ) /  P (4.34)

h4  h3  v3 ( P4  P3 ) /  P (4.35)

h7  h6  v6 ( P7  P6 ) /  P (4.36)

Where, ηP is the pump’s isentropic efficiency.


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 81

The fraction z is determined from mass and energy balance over the open FWH assuming
that the feed-water leaves the heater as saturated liquid. This leads to:

(1  y )(h3  h2 )
z (4.37)
(h12  h2 )

Since the two streams entering the closed feed-water heater are at different pressures and
leave it unmixed, the temperature of the feed-water at point 5 can take any value between
T4 and T10. In the present analysis it is assumed that the feed-water leaves the closed FWH
at the saturation temperature corresponding to the reheater pressure (Prh), i.e. T5=T6
=Tsat@Prh. Once the value of h5 is known, the mass fraction y can be determined according
to the following relationship:

(h5  h4 )
y (4.38)
h10  h6   h5  h4 
The enthalpy at state 8 is also obtained by energy balance as:

h8  1  y h5  yh7 (4.39)

The following numerical example, which is adapted from Example 10.6 given by Cengel
and Boles [2], analyses the cycle’s performance and proves its advantages compared to
the cycle with a single feed-water heater considered in Example 4.3.

Example 4.4. Analysis of the regenerative Rankine cycle with two heaters
Consider a steam power plant that operates on an ideal reheat–regenerative Rankine cycle
with one open feed-water heater, one closed feed-water heater, and one reheater. Steam
enters the turbine at 15 MPa and 600°C and is condensed in the condenser at a pressure
of 10 kPa. Some steam is extracted from the turbine at 4 MPa for the closed feed-water
heater, and the remaining steam is reheated at the same pressure to 600°C. The extracted
steam is completely condensed in the heater and is pumped to 15 MPa before it mixes
with the feed-water at the same pressure. Steam for the open feed-water heater is
extracted from the low-pressure turbine at a pressure of 0.5 MPa. The feed-water is heated
in the closed-feed-water heater to the saturation temperature at the feed-water heater
pressure (i.e., 4 MPa).

Solution
Figure 4.15 shows the Excel sheet developed for this example and Figure 4.16 reveals
the formulae used in the sheet. The sheet can account for anisotropic compression and
expansion, but for this ideal case the isotropic efficiencies of the pumps and turbines are
assigned values of 100%. The formula window reveals the formula used for the
82 Mohamed M. El-Awad

calculation of h5. Note that the function Wath_Px(P_rh,0) is used to determine the value
of hf at the reheat pressure (P_rh).

Figure 4.15. Excel sheet for Example 4.4

Figure 4.16. Excel formulae used for Example 4.4

Figure 4.15 shows that the net specific work is 1397.186 kJ while the thermal efficiency
is 48.255%. Compared to the regenerative cycle with a single feedwater heater, the net
specific work increased by 10.91 % while the thermal efficiency increased by 5.93%.
Table 4.5 compares the values of the important cycle parameters obtained in the present
analyses with their corresponding values given by Cengel and Boles [2]. The figures in
the table show good agreements between the present enthalpy values and those of Cengel
and Boles [2] for all the fluid states, but the calculated value of the steam fraction, y,
deviates from the corresponding values given by Cengel and Boles [2] by about 3%. As
a result, the calculated value for the cycle’s efficiency deviates by 1.92% from that that
given by Cengel and Boles [2].
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 83

Table 4.5. Verification of Thermax results for Example 4.4


Parameter Cengel and Boles [2] Thermax Deviation (%)
h8 1089.8 1089.833 0.00303
h9 3583.1 3579.440 -0.10214
h10 3155.0 3184.78 0.94390
h11 3674.9 3674.033 -0.02360
h12 3014.8 3019.572 0.15827
h13 2335.7 2338.217 0.10778
y 0.1766 0.171 -3.36863
z 0.1306 0.131 0.57504
 0.492 0.483 -1.91890

4.5. The combined Brayton-Rankine cycle


The combined Brayton-Rankine cycle, or simply the combined-cycle (CC), captures the
exhaust gas energy and utilises it for steam generation in a heat-recovery steam generator
(HRSG). When expanded in a steam turbine, the generated steam produces additional
power so that the overall efficiency of the plant can be increased by 50 – 60 present.
Figure 4.17 shows a schematic diagram of the CC power plant and Figure 4.18 shows its
T-s diagram. Note that states 2s and 4s are the states after idealised isentropic
compression and expansion in the Brayton cycle, while the states 2 and 4 are the actual
corresponding states. Similarly, state 8s in the Rankine cycle is that after an idealised
isentropic expansion in the steam turbine, while state 8 is the actual corresponding state.

C.C.
3
2
1
G
Compressor Gas turbine
4
5

6 7
HRSG
G
Steam turbine
8
Condenser
9
Pump
Figure 4.17. A schematic diagram for a combined-cycle power plant
84 Mohamed M. El-Awad

T 3

2s 2
4
5 4s 7

1
6

9 8s 8

s
Figure 4.18. T-s diagram for the combined-cycle power plant

Air enters the compressor of the gas-turbine at state 1 where it is compressed to state 2
and heated in the combustion chamber to state 3. After expanding in the turbine, the
exhaust gas leaves the gas turbine at state 4 and then passes through the HRSG where
part of its energy is used to produce steam for the steam power-plant and eventually
leaves the heat exchanger at state 5. The feed-water of the steam-turbine system enter the
HRSG as compressed liquid at state (6) before being heated to become superheated steam
at state (7). It is then delivered to the steam turbine where it expands to the condenser
pressure at state (8). After leaving the condenser as saturated liquid water at state (9), the
feed-water is pumped by the first feed-water pump to state (6). Usually the temperature
of the gas at point 5 is not allowed to drop below 150oC in order to avoid condensation
of the water vapour in the gas that will lead to corrosion of the duct and chimney.

4.5.1. First-law analysis of the combined cycle


For steady-state operation and negligible heat transfer with the surroundings, the energy
rejected by the Brayton cycle equals that gained by the Rankine cycle. Neglecting friction
losses and changes in kinetic and potential energy, energy balance in the HRSG gives:

 s h7  h6   m
m  g h4  h5  (4.40)

 s are the mass flow rates of the gas-turbine exhaust and the steam,
 g and m
Where m
respectively. Knowing both the inlet and exit temperatures of the exhaust gas and the
steam, Equation (4.40) can be used to relate the mass flow rates in the Brayton and
Rankine cycles as follows:

 s h4  h5 
m
 (4.41)
 g h7  h6 
m
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 85

Given the total net power of the combined plant, W , in MW, the mass flow rate of the
gas in the gas-turbine cycle can be obtained from:

W  1000
g 
m (4.42)

h3  h4   h2  h1   m s h7  h8   h6  h9 
g
m
 g ), the corresponding mass
Having determined the mass flow rate in the gas-turbine ( m
 s ) can be calculated from Equation (4.40).
flow rate of steam in the steam-turbine ( m

Enthalpy values h2, h4 and h8 are determined as follows:

h2  h1  h2s  h1  /  c (4.43)

h4  h3  h3  h4s  gt (4.44)

h8  h7  h7  h8s  st (4.45)

Where ηc, ηgt, and ηst, refer to the isentropic (adiabatic) efficiencies of the air-compressor,
the gas-turbine, and the steam-turbine, respectively. The thermal efficiency for the
Brayton cycle alone (ηBrayton) is given by:

 Brayton  WGT  WC  / Qin (4.46)

Where Qin is the heat input and WGT the work produced by the gas-turbine and WC is the
work input to the compressor. The thermal efficiency of the combined plant is given by:

 combined 
WGT 
 WC   WST  W p  (4.47)
Qin

Where WST is the work produced by the steam-turbine plant and Wp is the work consumed
by the pump. In terms of fluid flow rates and enthalpies, the five quantities on the right-
hand sides of Equations (4.46) and (4.47) can be expressed as:

Qin  m g h3  h2  (4.48)

WGT  m g h3  h4  (4.49)

WC  m g h2  h1  (4.50)
86 Mohamed M. El-Awad

WST  m s h7  h8  (4.51)

W p  m s v9 P6  P9  (4.52)

The following example is based on Example 9.13 in Moran and Shapiro [1].

Example 4.5. Energy analysis of the combined Brayton-Rankine cycle


In a 45-MW combined-cycle system such as the one shown in Figure 4.17, air enters the
compressor at 300K and 100 kPa. The compressor pressure ratio is 12 and air leaves the
combustion chamber at 1400 K. The isentropic efficiencies of the compressor and gas-
turbine are 84% and 88%, respectively. The exhaust gases are used to produce steam in
the HRSG at 8 MPa and 500oC. The exhaust gases leave the HRSG at 400K. The
condenser pressure is 8 kPa. The isentropic efficiencies of the steam-turbine and pump
are 90% and 80%, respectively. Determine the mass flow rates in the gas-turbine and the
steam-turbine cycles and the overall thermal efficiency of the plant.

Solution
Figure 4.19 shows the Excel sheet developed for this example by using Thermax
functions. The top part of the sheet deals with the analysis of the Brayton cycle, while
the bottom part deals with that of the Rankine cycle.

Figure 4.19. Excel sheet developed for Example 4.5 using Thermax
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 87

The Brayton cycle is analysed by using property functions from the Gas-group of
Thermax as shown in Example 4.1, while the Rankine cycle is analysed by functions
from the Wat-group of Thermax as shown in Example 4.4. The development of this sheet
is left as an exercise to the reader (Exercise 4.6).

Table 4.6 compares the calculated values of enthalpies at various points in the cycle with
their corresponding values obtained by Moran and Shapiro [1]. As the figures indicate,
the difference between the values calculated by Thermax and those given by Moran and
Shapiro [1] are minor.

Table 4.6. Verification of Thermax results for energy analysis of the combined cycle
Moran and Deviation
Quantity Thermax
Shapiro [1] (%)
h1 300.19 299.845 -0.115
h2 669.79 668.797 -0.148
h3 1515.42 1515.758 0.022
h4 858.02 858.636 0.072
h5 400.98 401.000 0.005
h9 173.88 173.357 -0.301
h6 183.96 183.431 -0.287
h7 3138.30 3155.111 0.536
h8 2104.74 2113.800 0.431
m g 100.87 100.676 -0.192
m v 15.60 15.504 -0.615
 Brayton - 0.340 -
 CC - 0.528 -

The figure in Table 4.6 show that the thermal efficiency of the simple Brayton cycle is
only 34%, while that of the combined cycle reached 52.8%. It should also be mentioned
that the gas-turbine and steam-turbine systems in the combined cycle need not be simple
systems as shown in Figure 4.17 since both of them can adopt various measures that
improve their efficiencies. For example, the gas-turbine can have intercooling and
reheating and the Rankine cycle can have single or multiple feed-water heaters.

4.5.2. Second-law analysis of the combined cycle


First-law analyses, also called energy analyses, are useful for evaluating the overall
performance of energy-utilisation systems. The performance indicator used in such
analyses is the thermal efficiency in the case of power-producing systems or the
coefficient of performance (COP) in the case of refrigeration and heat-pump systems. By
88 Mohamed M. El-Awad

comparison, second-law or exergy analyses aim to identify the locations and magnitudes
of energy losses in the system. Therefore, exergy analyses are more useful than energy
analyses for designing more efficient energy systems. This section illustrates the
application of the method to the combined cycle in Example 4.5.

For open systems (control volumes), the specific flow exergy (εf) is defined as [1]:

V2
 f  h  h0   T0 s  s 0    gz (4.53)
2

Where T0 is the dead-state temperature and h0 and s0 are values of enthalpy and entropy
of the working fluid at the dead-state (T0, P0). The last two terms that represent
contributions from kinetic and potential energy are usually small compared to the other
terms and can be neglected as follows:

 f  h  h0   T0 s  s0  (4.54)

In general, the rate of exergy increase or decrease across any open system that involves
work transfer ( W ) and heat transfer ( Q ) with its surroundings is governed by [1]:

 T0 
dEcv
 Q j  W cv  P0 dVcv   m   m   E
dt   e e fe d
 1  (4.55)
dt  Tj  
i fi
j   i

Where E d refers to the rate of exergy destruction and suffices i and e refer to the inlet
and exit conditions of the fluid. The sign convention of heat-in (+) and work-out (+)
applies in Equation (4.55), which for steady, single-pass flow devices, reduces to:

 T 
E d  1  0 Q  W cv  m  fi   fe  (4.56)
 T 

Furthermore, for devices that involve only adiabatic process (e.g. the air-compressor, gas-
turbine, pump and steam turbine), the equation becomes:

E d  W cv  m  fi   fe  (4.57)

Alternatively, the rate of exergy destruction for these devices can obtained from the
following relationship:

E d  m T0 se  si  (4.58)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 89

Applying Equation (4.58) to the gas-turbine cycle:

E d ,compressor  m g T0 s 2  s1  (4.59)

E d ,airturbine  m g T0 s 4  s3  (4.60)
Similarly, for the steam-turbine cycle:

E d , pump  m s T0 s6  s9  (4.61)

E d ,steamturbine  m s T0 s8  s7  (4.62)

For the steam generator that does involves neither work nor heat-transfer, Equation (4.55)
becomes:

 d   m i fi   m 
o fe  g  5   4   m
=m  s  7   6  (4.63)
i e

The net exergy increase of the gas passing through the combustion chamber is:

E gain, c.c.  m g  f 3   f 2  (4.64)

The rate of exergy loss with the exhaust gas is given by:

E loss, exhaust  m g  f 5   f 1  (4.65)

Finally, the rate of exergy loss in the condenser is given by:

E loss, condenser  m s  f 8   f 9  (4.66)

Figure 4.20 shows the extensions needed for the Excel sheet used for energy analysis of
the combined cycle shown in Figure 4.19. The additional data required by exergy analysis
include: the reference temperature (T_0), the reference pressure (P_0), and the gas
constant for air (R_air). Based on these data, the sheet calculates entropy values of the
air and water at various locations and then determines the rates of exergy gains and exergy
destructions in the different system components from the relevant equations. Figure 4.21
reveals Thermax property functions used in these calculations and Table 4.7 compares
values of the key parameters in the exergy analysis to those given by Moran and Shapiro
[1].
90 Mohamed M. El-Awad

Figure 4.20. Extension to the Excel sheet of the combined cycle in Example 4.5
required for the exergy analysis

Figure 4.21. Thermax functions used in exergy analysis of the combined cycle

Table 4.7. Accuracy of Thermax results for exergy analysis of the combined cycle
Moran and Thermax Deviation
Shapiro [1] (%)
Net exergy loss in the exhaust gas at state 5 1.39 1.400 0.7194
Exergy loss in the condenser cooling water 1.41 1.389 -1.4894
Exergy destruction in air-turbine 3.42 3.425 0.1462
Exergy destruction in the compressor 2.83 2.836 0.2120
Exergy destruction in the steam turbine 1.71 1.711 0.0585
Exergy destruction in the pump 0.02 0.00 -100
Exergy destruction in the heat exchanger 3.68 3.660 -0.5435

In general, Table 4.7 shows very small differences (less than 1%) between the present
values and those obtained by Moran and Shapiro [1] except for the rate of exergy
destruction in the pump which has been neglected in the present analysis by specifying
s6 = s9. Moran and Shapiro [1] obtained the value shown on the table by accounting for
the small difference between s6 and s9,
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 91

Exergy analysis of the combined cycle indicates that the maximum rate of exergy
destruction occurs in the HRSG (3.660 MW) followed closely by the air-turbine (3.425
MW) and the air-compressor (2.836 MW). Exergy losses in the exhaust gas and the
cooling water of the condenser have similar magnitudes and each one amounts to around
1.4 MW.

4.6. Closure
Thermax property functions in the “Wat” group enable Excel to be used for
thermodynamic analyses of various power-generation cycles based on the Rankine cycle.
The chapter illustrated this capability by analysing the simple Rankine cycle, the Rankine
cycle with superheat and reheat, and the regenerative Rankine cycle with one feedwater
heater and with two heaters. These analyses illustrate the advantages of the various
improvements made to the simple Rankine cycle gained by increasing the net-work
output and the thermal efficiency. Comparisons of the results obtained with Thermax
functions with those given by Moran and Shapiro [1] and by Cengel and Boles [2] prove
the accuracy of the relevant Thermax property functions.

The chapter also illustrated the use of the add-in functions in the “Wat” and “Gas” groups
for energy and exergy analyses of the combined Brayton-Rankine cycle. Thermodynamic
analyses of power-generation cycles with the Excel-based platform can also be performed
by using other educational property add-ins such as those developed at the Mechanical
Engineering Department at the University of Alabama [4], the University of Virginia [5],
or the TPX add-in developed by Goodwin [6]. Although not demonstrated in the chapter,
all the analyses presented in this chapter can be performed by using the two add-ins
developed at the University of Alabama [4]; “Thermotables” and “Idealgas”.

References
[1] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley, & Sons. Inc. 2006
[2] Y. A. Cengel and M. A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 8th Edition, 2015
[3] M.M. El-Awad and M.S. Al-Saidi, Verification of Thermax functions for Excel
analysis of steam-injection gas turbine cycle, International Journal of Research in
Engineering and Technology (IJRET) Vol. 4, No. 1, 2016, 3-8
[4] The University of Alabama, Mechanical Engineering, Excel for Mechanical
Engineering project, Internet: http://www.me.ua.edu/excelinme/index.htm (Last
accessed July 19, 2018)
[5] The University of Virginia, Free Excel/VBA Spreadsheets for Thermodynamics:
Rankine, Brayton, Otto and Diesel Cycles, Internet:
http://www.faculty.virginia.edu/ribando/ modules/xls/Thermodynamics/ (Last
accessed July 1, 2019)
[6] D. G. Goodwin, "TPX: thermodynamic properties for Excel", Available at:
http://termodinamicaparaiq.blogspot.com/p/tpx_12.html (Last accessed July 11,
2019)
92 Mohamed M. El-Awad

Exercises
1. Using the Excel sheet developed for Example 4.1, study the effect of increasing the
boiler pressure (Pb) on the thermal efficiency and wetness fraction at the last stages
of the low-pressure turbine. Plot these two parameters for the following values of Pb:
7, 10, 15, 20 MPa.
2. By making suitable extensions to the Excel sheet developed for first-law analyses of
the simple ideal Rankine cycle in Example 4.1, develop an Excel sheet for second-
law analysis of the cycle. Verify your results by comparison with those obtained by
Cengel and Boles [2] in their EXAMPLE 10–7.
3. Using the Excel sheet developed for Example 4.2, study the effect of varying the
reheat pressure on the dryness fraction of steam at the last stages of the turbine (x6),
the net specific work, and thermal efficiency.
4. A steam power plant operates on an ideal reheat-regenerative Rankine cycle and has
a net power output of 80 MW. Steam enters the high-pressure turbine at 10 MPa and
550oC and leaves at 0.8 MPa. Some steam is extracted at this pressure to heat the
feed-water in an open feed-water heater. The rest of the steam is reheated to 500oC
and is expanded in the low-pressure turbine to the condenser pressure of 10 kPa.
Make the necessary modifications to the Excel sheet developed for Example 4.2 and
use the modified sheet to study the effect of the intermediate pressure on the net work
output and thermal efficiency.
5. Consider the steam power plant shown in Figure 4.P5 that operates with reheat and
one closed feed-water heater.

3 HP LP
Turbine Turbine
4
y 6
Steam
Generator 5
8 Condenser

11 7 2
1
10
9 Pump I
Pump II

Figure 4.P5

Steam enters the high-pressure turbine at 15 MPa and 600°C where it expands to a
pressure of 700 kPa after which it is returned to the steam generator house for
reheating to a temperature of 550oC. The steam then expands in the low-pressure
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 93

turbine and then condensed in the condenser at a pressure of 10 kPa. Some steam
leaves the low-pressure turbine at an intermediate pressure Pi = 0.3 MPa and enters
the closed feed-water heater. The extracted steam exits the feed-water heat as
saturated liquid at the same pressure.
Using the ideal regenerative Rankine cycle model, develop an Excel sheet to
determine: (i) the fraction of steam extracted from the turbine (y) and (ii) the thermal
efficiency of the cycle (η). Use the Excel sheet to study the effect of varying the
steam extraction pressure (P8) on y and η.
6. Develop the Excel sheet used to analyse the combined Brayton-Rankine cycle in
Example 4.5. Use the Excel sheet to study the effects of the temperature T5 on the
overall thermal efficiency of the plant ηcombined and mass flow rates of the fluids in
the two cycles, m  g and m s , by varying the value of T5 from 200 to 600K.
94 Mohamed M. El-Awad
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 95

Chapter 5

Analyses of simple and multi-stage


VCR systems
96 Mohamed M. El-Awad

The vapour-compression refrigeration (VCR) systems used in various residential,


commercial, and industrial applications are the main contributor to the ozone-layer
depletion problem and a major contributor to the global-warming problem. Therefore,
improving the performance of these systems is a critical matter. The performance of the
simple single-stage VCR system deteriorates when there is a large difference between the
evaporator and condenser temperatures. In this case, multi-stage compression VCR
systems that split the refrigeration cycle into low-pressure, intermediate-pressure, and
high-pressure stages become more effective. In order to get the best performance of these
multi-stage VCR systems, their stages have to be optimised. This chapter presents
optimisation analyses of two-stage and three-stage compression cycles using refrigerant
R134a. The results obtained by the property functions provided by Thermax are verified
against the relevant values obtained by using the add-in developed for R134a at the
University of Alabama (UA) [1] and against those given by Moran and Shapiro [2] and
Cengel and Boles [3].

5.1. The ideal vapour-compression refrigeration cycle


Figure 5.1 shows a schematic diagram of the basic VCR system and Figure 5.2 shows the
T-s diagram for the ideal VCR cycle that consists of the following four processes:

1. Process 1-2: reversible adiabatic compression process


2. Process 2-3: constant-pressure heat-rejection process in the condenser
3. Process 3-4: expansion in an expansion value or a capillary tube
4. Process 4-1: constant-pressure heat-addition process in the evaporator

qout

Condenser

3 2

win
Expansion
Compressor
valve

1
4
Evaporator

qin

Figure 5.1. Schematic diagram of the single-stage vapour-compression


refrigeration system
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 97

T
2

Tc 3

Te
4 1

s
Figure 5.2. T-s diagram for the ideal vapour-compression refrigeration cycle

The ideal cycle does not account for the irreversibility and pressure losses in the system
components and the compressor is assumed to be adiabatic and reversible. Therefore, the
compression process is isentropic as shown in Figure 5.2. Assuming steady operation of
all the system components and neglecting changes in kinetic and potential energy across
each component, the energy interactions with the surroundings in the different
components per unit mass flow rate of the refrigerant in the system are described by the
following relationships.

Compressor input work:

wc  h2  h1  (5.1)

Heat rejection in the condenser:

q out  h2  h3  (5.2)

Heat input in the evaporator:

q in  h1  h4  (5.3)

Adiabatic process in the expansion valve or capillary tube:

h 4  h3 (5.4)
98 Mohamed M. El-Awad

Three parameters measure the performance of the VCR system; which are its coefficient
of performance (COP), cooling capacity (CC) in ton of refrigeration, and the compressor
power ( W c ):

qin
COP  (5.5)
wc

60 m q in
CC  (5.6)
211

 h2  h1 
W c  m (5.7)

Where m is the mass flow rate of the refrigerant in the system in kg/s. The following
example, which is based on Example 10-1 in Moran and Shapiro [2], illustrates the use
of Thermax and the UA add-in “R134a” for analysing the cycle.

Example 5.1. Analysis of the ideal vapour-compression refrigeration cycle


An ideal VCR cycle thermally connected to a cold region at 0 oC and a warm region at
26oC uses refrigerant R134a as the working fluid. The refrigerant enters the compressor
as saturated vapour at 0oC and leaves the condenser as saturated liquid at 26oC. The mass
flow rate of the refrigerant is 0.08 kg/s. Determine (a) the compressor power, in kW, (b)
the refrigeration capacity, in tons, and (c) the coefficient of performance.

Solution
Figure 5.3 shows the Excel sheet developed for this example by using the functions
provided by the refrigerants group (Ref) of Thermax. The given data for the evaporator
and condenser temperatures and refrigerant mass flow rate are shown on the left side of
the sheet as T_e, T_c, and m_ref, respectively. The middle part of the sheet shows the
formulae used for determining the different enthalpy values using the relevant functions
of Thermax. For example, the enthalpy at state 1, h_1, is determined by using the function
Refh_Tx that returns the enthalpy of saturated vapour (hv) at a given temperature. The
entropy at this point (s_1) is found by using another function, Rehs_Tx, that determines
the entropy of saturated vapour at a given temperature (T) and quality (x).

The right-hand side of the sheet calculates the compressor power, w_c, according to
Equation (5.7), the cooling capacity, CC, according to Equation (5.6), and the coefficient
of performance (COP) according to Equation (5.5). Figure 5.4 shows a similar sheet that
was developed by using property functions from the R134a add-in provided by the
University of Alabama [1]. Note that this add-in supports two systems of units, but its
functions can only be used for refrigerant R-134a. Table 5.1 that compares the calculated
values for the key cycle parameters by the two add-ins shows a good agreement between
the add-in results with their corresponding values given by Moran and Shapiro [2].
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 99

Figure 5.3. Excel sheet developed for Example 5.1 using Thermax

Figure 5.4. Excel sheet developed for Example 5.1 using R134a

Table 5.1. Comparison of the ideal VCR cycle key parameters given by the two add-ins
with their values given by Moran and Shapiro [2]
Parameter Moran & Thermax UA Add-in
Shapiro [2] R134a
Compressor work [kW] 1.40 1.409 1.409
Refrigeration capacity (ton) 3.67 3.700 3.700
COP 9.24 9.237 9.235

5.2. The actual vapour-compression refrigeration cycle


In actual VCR cycles, all four processes are irreversible because of friction losses in the
compressor and heat-transfer with finite temperature differences in the evaporator and
condenser. Some degree of subcooling also occurs at the condenser discharge. These
modifications to the ideal cycle affect both the cooling capacity and COP of the VCR
system. Figure 5.5 shows the T-s diagram for the actual VCR cycle.
100 Mohamed M. El-Awad

T 2
2s

Tc 3

Te
4 1

Figure 5.5. T-s diagram for the actual VCR cycle

For simplification, it is assumed that saturated vapour refrigerant enters the compressor
where it is compressed to a pressure that matches the required condenser temperature.
The actual exit temperature at state 2 is higher than the corresponding temperature in the
ideal cycle at state 2s, while the actual exit temperature from the condenser at state 3 is
lower than the corresponding temperature in the ideal cycle, which is Tc. Referring to
Figure 5.5, the relationships that determine the specific compressor work and heat-
rejection in the condenser become:

wc  h2  h1  (5.8)

q out  h2  h3  (5.9)

The enthalpy at state 2 can be determined from:

h2  h1  h2 s  h1  /  c (5.10)

Where ηc is the isentropic efficiency of the compressor. The enthalpy at state 3 is


approximately taken as the enthalpy of the saturated refrigerant at the given temperature.
The following example, which illustrates the use of the Thermax and R134a add-ins for
analysing the actual VCR cycle, is based on Example 5-3 in Moran and Shapiro [2].

Example 5.2. Analysis of the actual vapour-compression refrigeration cycle


A VCR system thermally connected to a cold region at 0oC and a warm region at 26oC
uses refrigerant R134a as the working fluid. The refrigerant temperature in the evaporator
is -10oC, while the condenser pressure is 9 bars, which corresponds to a saturation
temperature of 36oC for R134a. The refrigerant exits the condenser as subcooled liquid
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 101

at 30oC. Analyse the cycle to determine its COP and cooling capacity taking the isentropic
efficiency of the compressor as 80%.

Solution
Figure 5.6 shows the Excel sheet developed for analysing the cycle with Thermax. Note
the additional entries in the data and calculations parts. In the data part, the condenser
pressure is entered instead of the condenser temperature. The data part also added the
subcooled fluid temperature, T_3, and the compressor efficiency (η_c). The calculations
part adds a new formula for determining the compressor discharge enthalpy, h_2. Also,
note that the enthalpy at state 3 is now evaluated at 30oC and not at the saturation
temperature corresponding to the condenser pressure, which is 36oC.

Figure 5.6. Excel sheet developed for the actual VCR cycle of Example 5.2

A similar Excel sheet was developed using the UA add-in for R134a. Table 5.2 compares
values of the three cycle parameters wc, CC, and COP as determined by the two add-ins
with their corresponding values given by Moran and Shapiro [2]. The figures in the table
show that the results obtained by the two add-ins are slightly different from those given
by Moran and Shapiro [2]. It should be noted that the work input in the actual cycle is
more than that of the ideal cycle of Example 5.1, while the cooling capacity is lower.
Therefore, the COP dropped from 9.235 in the ideal cycle to 3.83 in the actual cycle,
which is more realistic.

Table 5.2. Comparison of the actual VCR cycle key parameters with their values given
by Moran and Shapiro [2]
Parameter Moran & Thermax UA add-in
Shapiro [2] R134a
Compressor work [kW] 3.10 3.154 3.153
Refrigeration capacity (ton) 3.41 3.434 3.433
COP 3.86 3.828 3.830
102 Mohamed M. El-Awad

5.3. Optimisation analysis of the ideal two-stage compression cycle


A number of industrial refrigeration applications have very low evaporator temperatures.
In such cases, the two-stage compression refrigeration system shown in Figure 5.7
becomes more feasible. In this system, a flash-chamber is added at an intermediate
pressure. Part of the refrigerant evaporates during the flashing process and mixed with
the refrigerant leaving the low-pressure compressor. The mixture is then compressed to
the condenser pressure by the high-pressure compressor. The liquid in the flash chamber
is throttled to the evaporator pressure to cool the refrigerated space as it vaporizes in the
evaporator. Figure 5.8 shows the T-s diagram for the ideal cycle.

Condenser
5 4

9
6
3 HPC

7 2

8 1
Evaporator LPC

Figure 5.7. Schematic diagram for a two-stage compression VCR system

Figure 5.8. T-s diagram for a two-stage compression VCR system


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 103

5.3.1. The analytical model


The refrigerant is assumed to leave the evaporator as a saturated vapour and both
compression processes are assumed to be isentropic. The refrigerant leaves the condenser
as a saturated liquid and is throttled to a flash chamber pressure. Taking the mass flow
rate of the refrigerant in the condenser as m , the fractions of this flow rate that are
separated after the flash chamber as saturated vapour m v and saturated liquid m l are
given by:

 v  x6 m
m  (5.11)

 l  1  x6 m
m  (5.12)

Where x6 is the quality at state 6. Therefore, the rates of heat removal in the evaporator
Q and heat rejection in the condenser Q are given by:
e c

Qe  m l h1  h8  (5.13)

Q c  m h4  h5  (5.14)

The compression work W c which has two parts one in the low-temperature cycle and
another in the high-temperature cycle, is given by.

W c  m l h2  h1   m h4  h9  (5.15)

The COP of the system is then given by:

Q e 1  x 6 h1  h8 
COP   (5.16)
Wc 1  x 6 h2  h1   h4  h9 

5.3.2. The optimisation analysis


Cengel and Boles [3] considered a two-stage VCR system with refrigerant R134a as the
working fluid. In what follows, their data and analysis results will be used to verify the
results obtained by using Excel and property functions provided by Thermax and the add-
in developed at the University of Alabama (UA) for refrigerant R134a [1].

Example 5.3. Optimisation of the ideal two-stage compression cycle


An ideal two-stage compression VCR system using refrigerant R134a as the working
fluid operates between evaporator and condenser pressures of 0.14 and 0.8 MPa,
respectively, and its flash chamber has a pressure of 0.32 MPa.
104 Mohamed M. El-Awad

Solution
Figure 5.9 shows the Excel sheet developed for analysing the two-stage compression
cycle using property functions. The three values of the evaporator, condenser, and flash-
chamber pressures are stored in the data part that occupies the left side of the sheet.
Calculations of the enthalpy values and the quality at state 6 are done in the two central
columns using Thermax functions. Based on the calculated values of state enthalpies, the
sheet determines the amount of heat absorbed in the evaporator (q_E), the heat rejected
in the condenser (q_C), the total compression work (w_comp), and the cycle’s COP
(COP). Figure 5.10 shows the formulae entered in the three columns that calculate the
intermediate and final results and Thermax functions they use. In addition to the three
Thermax functions shown in Table 5.1, this analysis requires the function
“Refs_Ph(“R134a”,P_flash,h_9)” to determine the entropy s9 from the known pressure
and enthalpy values at state 9.

Figure 5.9. The Excel sheet developed for the two-stage compression VCR system

Figure 5.10. Excel formulae and Thermax functions used in the Excel sheet

A similar Excel sheet was also developed using the relevant property functions provided
by the add-in developed at the University of Alabama for refrigerant R134a [1]. Table
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 105

5.4 compares the values of various cycle parameters obtained by the two Excel sheets
with their corresponding values given by Cengel and Boles [3]. The figures on the table
show that the values of Q , W , and COP obtained by both add-ins are very close to those
e c
given by Cengel and Boles [3].

Table 5.4. Analysis results for the two-stage compression cycle


Cengel and Thermax UA add-in Error (%)
Boles [3] R134a
h1 239.16 387.3095 387.3244 -0.0038
s1 1.740259 1.740194 0.0037
h2 255.93 404.0627 404.0133 0.0122
h3 251.88 400.0385 400.0513 -0.0032
h4 274.48 422.611 422.6194 -0.0020
h5 95.47 243.6307 243.6611 -0.0125
h6 95.47 243.6307 243.6611 -0.0125
x6 0.2049 0.204904 0.204879 0.0122
h7 55.16 203.3228 203.364 -0.0203
h8 55.16 203.3228 203.364 -0.0203
h9 255.10 403.2381 403.2016 0.0091
s9 1.737308 1.737218 0.0052
Q e 146.3 146.287 146.2708 0.0111
W c 32.71 32.69327 32.68758 0.0174
COP 4.47 4.47453 4.474811 -0.0063

The COP of the two-stage compression cycle can be improved by suitably adjusting the
pressure at the flash-chamber. The results obtained by changing the value of this pressure
are shown in Figure 5.11. The figure shows that the cycle’s COP can reach 5.49 at a flash-
chamber pressure of around 400 kPa. The maximum COP and corresponding flash-
chamber pressure can be determined more precisely by using Solver and Figure 5.12
shows the completed Solver Parameters dialog box for this task. Two constraints have
been imposed so as to keep the value of the flash-chamber pressure, P_flash, between the
specified values of the evaporator and condenser pressures. Figure 5.13 shows that the
solution reached by Solver by using the GRG Nonlinear method. According to this
solution the maximum COP is 4.49 and occurs at a flash-chamber pressure of 375.45 kPa.
Further improvement of the system’s COP is still possible by using more than one flash
chamber at different pressure levels as show in the following section.
106 Mohamed M. El-Awad

4.6

4.5

COP 4.4

4.3

4.2

4.1
0 200 400 600 800
Flach-chamber pressure (kPa)
Figure 5.11. Variation of the COP with the flash-chamber pressure in the two-stage
compression VCR system

Figure 5.12. Solver Parameters dialog box for optimisation of the two-stage
compression cycle

5.4. Optimisation analysis of the three-stage compression cycle


By using two flash chambers instead of one chamber, three-stage compression VCR
systems can be more cost-effective than two-stage compression systems if the
temperature lift is sufficiently high. Figure 5.14 shows a schematic diagram for such a
system and Figure 5.15 shows its ideal T-s diagram. Depending on the refrigerant used
in the system, there are certain values of the two flash-chamber pressures that maximise
the system’s COP. However, unlike the two-stage compression system, the determination
of these pressures requires the use of an optimisation software since it is not as easy to
find them by varying the flash-chamber pressures as in the case of a single flash chamber.
In what follows, the optimum flash-chamber pressures will be determined by using
Solver.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 107

Figure 5.13. Solver solution for optimisation of the two-stage compression cycle

Condenser 11
12

13 10 HPC
14

5 4

9 IPC
6 3

7 2

LPC
8 1
Evaporator

Figure 5.14. Schematic of the three-stage compression VCR system

5.4.1. The analytical model


Taking the mass flow rate of the refrigerant in the condenser to be m , the mass flow rates
of the vapour and liquid fractions after the high-pressure flash chamber are given by:

m 14  x13 m
(5.17)

m 5  1  x13 m
(5.18)
108 Mohamed M. El-Awad

Figure 5.15. T-s diagram for the ideal three-stage compression VCR cycle

Accordingly, the mass flow rates of the vapour and liquid fractions after the low-
pressure flash chamber are given by:

m 3  x6 1  x13 m (5.19)

 7  1  x6 1  x13 m
m 
(5.20)

The total compression work and the rate of heat rejection in the condenser are now given
by:

Wct  m 7 h2  h1   m 3  m 7 h4  h9   m h11  h10  (5.21)

Q c  m h11  h12  (5.22)

Therefore, the system’s COP is given by:

COP  Q c / W ct (5.23)

5.4.2. The optimisation analysis


The advantage of the three-stage compression cycle over the simple and the two-stage
compression cycles is illustrated by the following example in which the evaporator and
condenser pressures in the three systems are identical.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 109

Example 5.4. Optimisation of the ideal three-stage compression cycle


The system to be analysed has the same evaporator and condenser pressures as in the
two-stage cycle considered earlier, i.e., 0.14 and 0.8 MPa, respectively. It is required to
determine values of the two flash-chamber pressures that maximise the system’s COP.

Solution
Figure 5.16 shows the Excel sheet prepared for this example by extending that of the two-
stage compression cycle shown in Figure 5.7. The data part now includes two flash-
chamber pressures, P_fc1 and P_fc2 instead of one. The two flash-chamber pressures are
initially assigned values of 320 kPa and 520 kPa. A third column has also been added to
the calculations part to determine the enthalpy and entropy values in the high-pressure
stage. The formulae are not shown, but left as an exercise for the reader (Exercise 5.6)

Figure 5.16. Excel sheet for analysing the three-stage compression VCR cycle

Figure 5.16 shows that at the specified values of the pressures at the two flash chambers,
the COP is 4.65. Although this COP is higher than the optimum pressure obtained earlier
for the two-stage compression cycle, which is 4.49, it is still not the maximum possible
value for the COP of the three-stage cycle. The maximum COP can be determined by
optimising the two flash-chamber pressures by using Solver. Solver Parameters dialog
box is show in Figure 5.17. As Figure 5.17 shows, three constraints have been imposed
so as to keep the value of the two flash-chamber pressures, P_fc1 and P_fc2, between the
specified values of the evaporator pressure P_evap, and the condenser pressure, P_cond,
i.e.:

P_evap < P_fc1

P_fc1 < P_fc 2

P_fc2 < P_cond


110 Mohamed M. El-Awad

Figure 5.17. Solver Parameters box for the optimisation of the three-stage cycle

Figure 5.18 shows the solution determined by Solver according to which the maximum
COP is reached when the flash-chamber pressures are 283.465 kPa and 498.434 kPa.
Comparison of the cycle’s overall parameters shown in Figure 5.18 with those shown in
Figure 5.13 indicates that the cooling rate per unit flow rate of the refrigerant (q_E) has
increased from 146.388 kJ for the two-stage system to 147.158 kJ for the three-stage
system, while the corresponding compressors work decreased from 32.595 kJ to 31.579
kJ. As a result, the COP of the three-stage system reached 4.66; which is higher than that
of the two-stage compression cycle (COP=4.49) by about 3.23% and higher than that of
the simple VCR cycle (COP=3.97) by about 14.44%. The advantage of the three-stage
system over the simple and the two-stage compression systems becomes more profound
as the temperature lift increases.

Figure 5.18. Solution determined by Solver for the three-stage compression VCR cycle
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 111

5.5. Closure
This chapter illustrated the use of Thermax functions for the analyses of basic
refrigeration cycles including the ideal VCR cycle and the actual VCR cycle. The results
obtained by using Thermax functions for refrigerant R134a were verified with those
given in standard textbooks or determined by another property add-in developed at the
Mechanical Engineering Department at the University of Alabama [1]. The chapter also
showed how the Excel-based modelling platform can be used for analysing and
optimising two-stage and three-stage compression VCR systems.

As mentioned in Chapter 2, Thermax functions that calculate the enthalpy and entropy of
superheated refrigerants are suitable only for subcritical pressures and temperatures.
Therefore, these functions can be used to analyse the effects of various improvements to
the simple VCR cycle, such as adding a suction-line heat exchanger to the system [4],
but they may not suit certain applications that require supercritical pressures such as
supercritical organic Rankine cycles.

References
[1] The University of Alabama, Website of Excel in Mechanical Engineering,
https://www.me.ua.edu/ExcelinME/
[2] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley & Sons, Inc, 2006.
[3] A. Cengel, and M.A. Boles, Thermodynamics an Engineering Approach, McGraw-
Hill, 8th Edition, 2015
[4] M.M. El-Awad, Effect of Suction-Line Heat Exchangers on the Performance of
Alternative Refrigerants to R-22, IOSR Journal of Mechanical and Civil
Engineering (IOSR-JMCE), Volume 13, Issue 3 Ver. IV (May- Jun. 2016), PP 109-
117

Exercises
1. It is required to compare the performance of refrigerant R134a as an alternative fluid
to refrigerant R22 for air-conditioning (air-cooling) applications. By suitably
modifying the Excel sheet developed for Example 5.1, compare the compressor
power (kW), refrigeration capacity (ton), and coefficient of performance of the
vapour-compression refrigeration cycle with both refrigerants for an evaporator
temperature of 10oC and various condenser temperatures typical to your
environment.
2. It is required to compare the performance of refrigerant R134a as an alternative fluid
to refrigerant R22 for air-conditioning (air-heating) applications. By suitably
modifying the Excel sheet developed for Example 5.1, compare the compressor
power (kW), heating capacity (kW), and coefficient of performance of the vapour-
compression refrigeration cycle with both refrigerants for a condenser temperature
of 40oC and various evaporator temperatures typical to your environment.
3. Modify the Excel sheet developed for Example 5.2 to model the performance of the
actual VCR cycle with R22 as refrigerant. Take the evaporator temperature as -10oC,
112 Mohamed M. El-Awad

the condenser pressure as 1390 kPa (at which the saturation temperature is 36 oC),
the compressor exit temperature as 30oC, and the compressor isentropic efficiency
as 80%.
4. Repeat Exercise 5.3 with R12 or R22 as the refrigerant using suitable evaporator and
condenser pressures.
5. By making suitable extensions to the Excel sheet developed for Example 5.2,
determine the rates of exergy destruction within the compressor and expansion
valve, in kW, for T0 = 299K (26oC). (Answer: 0.58 kW and 0.39 kW).
6. Develop the Excel sheet used in Example 5.4 to analyse the three-stage CVR cycle
with refrigerant R134a. Use the Excel sheet to repeat the analysis for two other
refrigerants that are considered suitable substitutes to refrigerant R134a. Discuss
your results.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 113

Chapter 6

Analyses of cascade VCR cycles


114 Mohamed M. El-Awad

Like multi-stage compression vapour-compression refrigeration (VCR) systems, cascade


systems are more effective than the simple system by splitting the refrigeration cycle into
low-pressure and high-pressure circuits. Cascade VCR systems are most advantageous
when the temperature lift between the evaporator and condenser is high enough to justify
using different refrigerants for the low-temperature and the high-temperature circuits.
However, these systems are more expensive and, in order to get their best performance,
the intermediate pressure(s) between the cycle stages have to be optimised. Using
Thermax functions for refrigerants properties, this chapter first analyses the ideal cycle
with a single fluid (R134a) and shows how Solver can be used for optimisation analyses
of cascade systems. The actual cycle is then analysed with three fluid pairs using R744
in the low-pressure stage and R134a, R410A, and R717 in the high-pressure stage.
Thermax results are compared with those given by Cengel and Boles [1], Messineo and
Panno [2], and Parekh and Tailor [3].

6.1. The analytical model for the cascade refrigeration cycle


Figure 6.1 shows a schematic diagram for a two-stage cascade refrigeration system. The
two different refrigerants are selected such that each one of them suits the low-pressure
circuit or the high-pressure circuit better than the other. The T-s diagram of the cycle is
shown on Figure 6.2.

7 Condenser A
6

Circuit A
8 Evaporator A HP
5 Compressor
3
Condenser B 2
Circuit B
4 LP
1 Compressor
Evaporator B

Figure 6.1. Schematic for a two-stage cascade refrigeration system

The coefficient of performance (COP) of the cascade refrigeration system depends on the
refrigerants used in the two circuits, the evaporator and condenser temperatures, the
temperature difference ΔT between the two refrigerants in the heat-exchanger, and the
adiabatic efficiency of the compressors. The effects of these factors on the system’s COP
can be assessed by developing a thermodynamic analytical model for the system.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 115

∆T

Figure 6.2. T-s diagram for a two-stage cascade refrigeration system

To simplify the model, the refrigerants leaving the two evaporators are assumed to be
saturated vapours as shown on T-s diagram of Figure 6.2. Given the condenser pressures
or temperatures in circuit A and circuit B, the ideal enthalpies following the two
compression processes, i.e. h2 and h6, are determined based on the requirement that s2 =
s1, and s6 = s5. The actual enthalpies h2a and h6a are then determined from:

h2 a  h1  h2  h1  /  c (6.1)

h6 a  h5  h6  h5  /  c (6.2)

Where ηc is the isentropic efficiency of the compressors (assumed to be the same).

Suppose that the mass flow rate of the refrigerant in the evaporator of circuit B ( m B ) is
known and that the losses of the heat-exchanger to the surroundings are negligible. Then
the mass flow rate in the condenser of circuit A ( m A ) is given by:

m A  m B h2 a  h3  / h5  h8  (6.3)

The rate of heat removal in the high and low-pressure evaporators, Q eA and Q eB ,
respectively, are given by:

Q eA  m A h5  h8  (6.4)

Q eB  m B h1  h4  (6.5)
116 Mohamed M. El-Awad

The enthalpies h4 and h8 after the adiabatic throttling processes are determined from:

h4  h3 (6.6)

h8  h7 (6.7)

The compression work in the high-pressure circuit, W cA , and the low-pressure circuit,
W cB , are given by:

 A h6a  h5 
W cA  m (6.8)

 B h2a  h1 
W cB  m (6.9)

The coefficient of performance of the high-pressure circuit COPA and the low-pressure
circuit COPB are given by:

COPA  Q eA / W cA (6.10)

COPB  Q eB / W cB (6.11)

Finally, the total compression work W c and overall COP of the cascade system are given
by:

W c  W cA  W cB (6.12)

COP  Q eB / W c (6.13)

It should be noted that six out of eight enthalpy values required in the above model are
those of the refrigerants as saturated liquid or saturated vapour, viz. h1, h3, and h4 in circuit
B and h5, h 7, and h8 in circuit A. As mentioned in Chapter 2, Thermax functions for
saturated refrigerants precisely determine the refrigerant enthalpies from ASHRAE data.
Only the two enthalpy values h2 and h6 have to be handled by Themax functions for the
superheated-region. Although the functions for determining the enthalpy of superheated
refrigerants are less accurate than those of saturated refrigerants, Chapter 5 showed that
their effect on the overall analysis is minor. The following subsection analyses the
performance of a cascade system using refrigerant R134a in both circuits.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 117

6.2. Optimisation of the ideal cascade refrigeration cycle using R134a


The COP of a cascade system can be higher than that of a single-stage one even if the
same refrigerant is used in both stages. This is demonstrated by the following example
which is based on Example 11-4 in Cengel and Boles [1].

Example 6.1. Optimisation of the cascade cycle using R134a


Consider a cascade refrigeration system that operates with refrigerant R134a as the
working fluid in both the low and high circuits between the pressure limits of 0.14 and
0.8 MPa. Assuming an ideal cycle in which ηc = 100% and the temperature difference in
the heat-exchanger is zero, determine the optimum pressure for the heat exchanger.

Solution
Figure 6.3 shows the Excel sheet developed for the ideal cascade CVR cycle. The given
data that include the two refrigerant used and the three values of the pressures are stored
in the data part that occupies the left side of the sheet. Note that the refrigerant names are
given as input arguments so that the same sheet can be used for various refrigerant pairs
by simply adjusting the refrigerant names in the data part. In the present case, the
refrigerants in both cycles, RefA and RefB, are specified as R134a. Figure 6.4 shows the
formulae entered in the three columns that calculate the intermediate and final results.

Figure 6.3. The Excel sheet developed for the ideal cascade VCR system

Figure 6.4. Excel formulae and add-in functions used in the Excel sheet
118 Mohamed M. El-Awad

The value given for the heat-exchanger pressure (P_HX=320 kPa) is that used by Cengel
and Boles [1]. Calculations of the enthalpy values are done in the two central columns
using property functions. Figure 6.4 shows the formulae entered in these two columns
and the property functions used in them. Note that only three Thermax function are used
in the Excel sheet since most of the enthalpies appearing in above analytical model are
those of saturated liquid or saturated vapour at the given pressure. The required input
parameters of the three functions are shown in Table 6.1.

Table 6.1. Thermax functions used in the modelling the cascade VCR system
Function Purpose and input parameters
Refh_Px(Ref,P,x) determines the enthalpy of a saturated refrigernat given its
pressure and quality (used to determine h1, h3, h5, and h7)
Refs_Px(Ref,P,x) determines the entropy of a saturated refrigernat given its
pressure and quality (used to determine s1, and s9)
Refh_Ps(Ref,P,s) determines the enthalpy of a sperheated refrigernat given its
pressure and entropy (used to determine h2)

Based on the calculated values, the sheet determines the mass flow rate and the amount
of heat absorbed in the evaporator of cycle B (m_B and q_L), the heat rejected in the
condenser of cycle A (q_H), the total compression work (w_comp), and the cycle’s
coefficient of performance (COP). A second Excel sheet was also prepared by using the
add-in developed at the University of Alabama for refrigerant R134a [4]. Table 6.2
compares the values of various cycle parameters obtained by the two Excel sheets with
the corresponding values given by Cengel and Boles [1].

Table 6.2. Verification of Thermax functions for the ideal cascade cycle
Cengel and Boles [1] Thermax UA add-in Error (%)
h1 239.16 387.3095 387.3244 0.003847
s1 1.740259 1.740194 -0.00374
h2 255.93 404.0627 404.0133 -0.01223
h3 55.16 203.3228 203.364 0.020259
h5 251.88 400.0385 400.0513 0.0032
s5 1.725766 1.725785 0.001101
h6 270.92 419.041 419.071 0.007159
h7 95.47 243.6307 243.6611 0.012476
mB 0.039 0.038958 0.038971 0.033358
Q eB 7.18 7.167723 7.169126 0.01957
W c 1.61 1.602793 1.601369 -0.08892
COP 4.46 4.472021 4.476872 0.108357
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 119

The figures in Table 6.2 show that enthalpy and entropy values determined by the two
add-ins are almost identical even for states 2 and 6 that lie in the superheated region.
Although the enthalpy values given by the two add-ins are different from those given by
Cengel and Boles [1] because the latter uses different reference values for enthalpy and
entropy, the values of Q , W , and COP given by both add-ins are very close to those
e c
given by Cengel and Boles [1]. The results obtained by changing the heat-exchanger
pressure are shown in Figure 6.5. The figure shows that the COP of the cascade cycle can
reach 4.48 at a heat-exchanger pressure of around 400 kPa. This COP is significantly
higher than that of the simple cycle working under the same evaporator and condenser
pressures, which is 3.97 [1]. The maximum COP and corresponding heat-exchanger
pressure can be determined more precisely by using Solver. Figure 6.6 shows Solver
Parameters dialog box for maximising the COP calculated in cell K8 by adjusting the
heat-exchanger pressure (P_HX) stored in cell K10 as shown in Figure 6.3.
4.6

4.5

4.4
COP

4.3

4.2

4.1
0 200 400 600 800
Pressure (kPa)
Figure 6.5. Variation of the COP with the heat-exchanger pressure

Figure 6.6. Solver’s parameters box for Example 6.1


120 Mohamed M. El-Awad

Two constraints have been imposed on the iterative solution so as to keep the value of
the heat-exchanger pressure, between the specified values of the evaporator and
condenser pressures. The solution found by Solver is shown in Figure 6.7. According to
this solution, the maximum COP is 4.49 which achieved at a heat-exchanger pressure of
374.9 kPa.

Figure 6.7. Solver’s solution for Example 6.1

6.3. Analysis of the actual cascade cycle with R744-refrigerant pairs


Messineo and Panno [2] used EES [5] to analyse the performance of a cascade VCR
system with carbon dioxide (R744) as the refrigerant in the low-temperature circuit and
various refrigerants in the high-temperature circuit. The following example is based on
their analyses, but uses Thermax functions to analyse the cycle with R744 in the low-
temperature circuit and R134a, R404A, R410A, or R290 in the high-temperature circuit.

Example 6.2. Analysis of the cascade cycle with R744-refrigerant pairs


A cascade refrigeration system operates with R744 in the low-temperature circuit with a
mass flow rate of 0.2 kg/minute. Compare the performance of the performance of the
optimised systems with R134a, R410A, and R717 as refrigerants in the high-temperature
circuit. The reference temperatures of the system are TeB = -35oC, TcA = 35oC, and ΔT =
5oC. Take the isentropic efficiency of both compressors as 70%.

Solution
Figure 6.8 shows the Excel sheet prepared for this example with R314a as the high-
temperature refrigerant. The data part on the left side of the sheet shows the refrigerants
pair used in the analysis (R134a and R744), the temperatures of the evaporator (T_eB),
the condenser (T_cA), and heat-exchanger (T_cB), the temperature difference in the heat-
exchanger (∆T), the mass flow rate of R744 in circuit B (m_B), and the compressors’
efficiency (η_c). Figure 6.9 shows the formulae used in the calculation part. The key
cycle parameters calculated in right side of the sheet include the mass-flow rate in circuit
A (m_A), the flow-rate ratio (mr) as given by Equation (6.3), the compressor work in
cycles A and B (w_cA and w_cB), the heat rejection in cycle A (q_CA), the total
compression work (w_comp), and the overall COP (COP).
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 121

Figure 6.8. Excel sheet developed for Example 6.2

Figure 6.9. Excel formula and Thermax functions used in the calculations part of the
Excel sheet of Example 6.2

Note that the refrigerant names in Thermax functions are given as input arguments so
that the same sheet can be used to analyse the performance of the cascade system with
R410A or R717 in the high-temperature circuit by changing name of the fluid in cell B2
to “R410A” or “R717” instead of “R134a”. For a specified heat-exchanger temperature
of 0oC, Figure 6.8 shows that the cycle’s COP is 1.639 and the ratio of the mass-flow rate
of R134a to that of R744 is 2.055. By changing the heat-exchanger temperature from -
25oC to 15oC, the cycle’s key COP and mass-flow ratio were obtained for the three
refrigerant pairs and plotted in Figures 6.10 and Figure 6.11, respectively.

Figure 6.10 compares the results obtained by the present add-in for the COP with those
obtained by Messineo and Panno [2] under the same conditions. The figure shows that
the results obtained by Thermax agree well with those obtained by Messineo and Panno
[2] by using EES. The highest COP of the cycle is obtained with the R744-R717 pair and
the lowest is that obtained with R744-R410A pair. The figure also shows that the
temperatures that maximises the COP of the cascade system for the three refrigerants
pairs are different.
122 Mohamed M. El-Awad

1.75

1.65

COP
1.55

1.45

1.35
-30 -20 -10 0 10 20
TcB
R744-R717
R744-R290
R744-R410A
(a) (b)
Figure 6.10. Effect of low-temperature stage condensing temperature on COP,
(a) results obtained by Messineo and Panno [2], (b) present results

3.5
3
2.5
2
COP

1.5
1
0.5
0
-30 -20 -10 0 10 20
TcB
R744-R717
R744-R290
R744-R410A
(a) (b)
Figure 6.11, Effect of low-temperature stage condensing temperature on mass ratio,
(a) results obtained by Messineo and Panno [2], (b) present results

Figure 6.11 shows the effect of the heat-exchanging temperature on the ratio of the two
mass-flow rates in cycle A and cycle B compared to those obtained by Messineo and
Panno [2]. This figure also shows good agreement between the results obtained by
Thermax and those determined by EES for all three refrigerant pairs. In general, the
lowest mass flow-rate ratio is that obtained by the R744-R717 pair and the highest is that
for R744-R134a pair.

The key cycle parameters with optimised heat-exchanger temperatures for the three pairs
of refrigerants were obtained by using Solver and the results are shown in Table 6.3. Note
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 123

that the optimised value of the COP for the cycle with R744-R717 is significantly higher
than those with R744-R134a and R744-R410A, while the mass flow rate is significantly
lower. The figures in the table also show that both the compressor work and heat rejected
in the high-temperature condenser for the cycle with R744-R717 are higher than their
corresponding values for the cycles with R744-R134a and R744-R410A.

Table 6.3. Optimised cycle parameters for the cascade cycle with three pairs
THX oC COP mr WcA [kW] qcA [kW]
R744-R717 -13.0 1.728 0.287 0.375 1.402
R744-R134a -7.82 1.658 2.155 0.342 1.361
R744-R410A -6.26 1.604 1.903 0.342 1.359

6.4. Analysis of the cascade cycle with refrigerants R507A and R23
A number of researchers used computer-based models to analyse the performance of
potential environment-friendly refrigerants for cascade VCR systems. In this respect,
Parekh and Tailor [3] conducted a thermodynamic analysis of the cascade system using
two ozone friendly refrigerants; R507A in the high-temperature circuit and R23 in the
low-temperature circuit. They took the mass flow rate in the low-temperature circuit, m B ,
as 0.2 kg/min and used an adiabatic efficiency of 0.8 for both compressors. They adopted
the following standard operating parameters:

1. Evaporator temperature in the low-temperature cycle (TeB) = -80°C


2. Condenser temperature in the high-temperature cycle (TcA) = 35°C
3. Condenser temperature in the low-temperature cycle (TcB) = -36°C
4. Temperature difference in cascade condenser (∆T) = 2.5°C

By using EES to develop their computational model, Parekh and Tailor [3] analysed the
system’s performance under different operating conditions. In one part of their study,
they varied the evaporator temperature in cycle B from -80°C to -60°C, the condenser
temperature in cycle A from 25°C to 50°C, and the condenser temperature in cycle B
from -50oC to 0oC. This part of their analysis was reproduced by using the present add-
in.

Figure 6.12 shows the Excel sheet prepared for this purpose based on the analytical model
described in the previous section. The data part on the left side of the sheet shows the
given information for the two refrigerants in the high-temperature and the low-
temperature circuits (RefA and RefB), the three temperatures in the low-temperature
evaporator (T_eB), the high-temperature condenser (T_cA), and low-temperature
condenser (T_cB). The data part also includes the temperature difference in the heat-
exchanger (ΔT), the mass flow rate in the low-temperature cycle (m_B), and the
compressors adiabatic efficiency (η_c).
124 Mohamed M. El-Awad

Figure 6.12. Excel sheet prepared for the analysis of the cascade system with R507A
and R23

Based on the given data, the sheet determines the high-temperature cycle’s evaporator
temperature (T_eA), and the pressures of the condensers in the high-temperature and low-
temperature cycle (P_cA and P_cB). Enthalpy values at the eight states in the upper and
lower circuits are calculated in the two central columns of the sheet. The last column to
the right shows the calculated mass flow rate in circuit A (m_A), the mass-flow ratio
(mr), and the rates of heat and work inputs in the low and high-temperature circuits and
then the total work input and COP of the cascade system. The sheet was used to determine
the overall COP, the amount of heat removed in the low-temperature evaporator and the
total compressor work by varying the evaporator temperature in circuit B (T_eB) from -
80oC to -50oC while keeping the two other temperatures T_cA and T_cB at their reference
values of 35.0oC and -36oC, respectively. Figure 6.13 compares the results obtained with
those given by Parekh and Tailor [3]. The value of the COP obtained by the present add-
in at T_eB = -50oC, which is 1.2296, compares well with its value of 1.232 obtained by
Parekh and Tailor [3].

1.3 800

1.2 750

1.1 700
Wc, QeB (W)
COP

1 650

0.9 600

0.8 550

0.7 500
-80 -70 -60 -50
TeB(oC)
(a) (b)
Figure 6.13. Effect of the low-temperature cycle evaporator temperature on COP, QeB
and Wc: (a) adapted from Parekh and Tailor [3], (b) present results
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 125

Figure 6.14 shows the results obtained by the sheet for different values of the condenser
temperature in cycle A (T_cA) from 25 to 50oC while fixing TeB and TcB at their
reference values and Figure 6.15 shows the results obtained by varying the condenser
temperature in cycle B (T_cB) from -50 to 0oC while keeping T_eB and T_cA at their
reference values. Clearly, both figures show a good agreement with the results obtained
by Parekh and Tailor [3]. Figure 6.15 also shows that the value of the condenser
temperature in the low-temperature circuit that maximises the COP is well estimated by
the present model.

0.95 1100
0.9
1000
0.85

Wc, QeB (W)


0.8 900
0.75

COP
800
0.7
0.65 700
0.6
600
0.55
0.5 500
25 30 35 40 45 50
TcA(oC)
(a) (b)
Figure 6.14. Effect of the high-temperature stage condensing temperature on COP, QeB
and Wc: (a) adapted from Parekh and Tailor [3], (b) present results

0.8 4.5

0.78 4
3.5
0.76

COPA, COPB
3
COP

0.74
2.5
0.72
2
0.7 1.5
0.68 1
-50 -40 -30 -20 -10 0
TcB(oC)
(a) (b)
Figure 6.15. Effect of the low-temperature stage condensing temperature on COP,
COPA, and COPB: (a) adapted from Parekh and Tailor [3], (b) present results

6.5. Closure
This chapter tests the adequacy of Thermax functions for analysing cascade VCR systems
with both synthetic and natural refrigerants. The cascade systems considered in the
analyses include a system with R134a in both the low-temperature and high-temperature
circuits, a system with R744 in the low-temperature circuit and R134a, R410A, and R717
126 Mohamed M. El-Awad

in the high-temperature circuit, and a system with R23 in the low-temperature circuit and
R507A in the high-temperature circuit. Comparison of the results obtained with those
obtained by EES software show good agreements.

Most of the fluid states in the analyses of VCR cycles lie in the compressed liquid or
saturated mixture regions and, therefore, the accuracy of Thermax functions for these
states is not questionable because they calculate the refrigerants properties using the data
provided by ASHRAE. Regarding the accuracy of estimating the fluid properties in the
superheated region by the relevant Thermax functions, the analyses presented in this
chapter showed that the accuracy of these functions has a minor effect on the accuracy of
estimating the overall cycle parameters.

A number of Excel add-ins have been developed by academic and research institutions
that can be used to widen the scope of analyses of VCR systems by using the Excel-based
platform. A good example of these add-ins is the REFPROP add-in developed by the
National Institute of Standards and Technology (NIST). REFPROP provides properties
of a large number of conventional and alternative refrigerants including refrigerants
mixtures [6]. A free-source alternative to REFPROP is the CoolProp software developed
by Bell [7].

References
[1] Y. A. Cengel and M.A. Boles. 2007, Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition.
[2] A. Messineo and D. Panno, 2012, Performance evaluation of cascade refrigeration
systems using different refrigerants, International Journal of Air-Conditioning and
Refrigeration Vol. 20, No. 3
[3] A. D. Parekh and P. R. Tailor , 2011, Thermodynamic Analysis of R507A-R23
Cascade Refrigeration System, World Academy of Science, Engineering and
Technology International Journal of Mechanical and Mechatronics Engineering
Vol:5, No:9
[4] The University of Alabama, Website of Excel in Mechanical Engineering,
https://www.me.ua.edu/ExcelinME/
[5] F-Chart Software, https://www.fchartsoftware.com/ees/
[6] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Reference Fluid Thermodynamic
and Transport Properties— REFPROP Version 8.0, User’s Guide, National Institute
of Standards and Technology, Physical and Chemical Properties Division, Boulder,
Colorado 80305, 2007.
[7] I. Bell, CoolProp. Available at https://sourceforge.net/projects/coolprop/files/
CoolProp/ (Last accessed July 11, 2019)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 127

Chapter 7

Analyses of air-conditioning processes


and systems
128 Mohamed M. El-Awad

Design analyses of air-conditioning systems require the thermodynamic properties of


humid air to be determined. The traditional method used in such analyses by visual
interpolation of psychrometric charts can be both time-consuming and inaccurate. This
chapter shows how Thermax functions can be used in such analyses for determining the
properties of air-water vapour mixtures. The chapter initially describes the basic
functions of the psychrometric chart before using Thermax functions for analysing two
basic air-conditioning processes; which are the cooling-with-dehumidification process
and the process of adiabatic mixing of two moist-air streams. Thermax functions are then
used to analyse two relevant systems, which are the evaporative cooler and the wet
cooling tower. Finally, the chapter shows how the property functions can be used for
solving a design optimisation problem that requires the energy consumption of an air-
conditioning system to be minimised by air re-circulation.

7.1. The psychrometric chart and the basic air-conditioning processes


The psychrometric chart is used in the analyses of air-conditioning systems and processes
to determine the state of humid air, usually at a given pressure in terms of its properties
at various temperatures and humidity levels. The pressure is usually the standard
atmospheric pressure of 1 atm (equivalent to 101.325 kPa). Figure 7.1 shows a typical
layout of the psychrometric chart.

Barometric pressure
Relative humidity (ϕ)

Enthalpy scale Wet-bulb


temperature Absolute humidity
(Twb) scale

Volume per
Enthalpy per kg dry air (v)
kg dry air (h)

Twb

Dry bulb temperature

Figure 7.1. The psychrometric chart

On the horizontal axis, the chart shows the dry-bulb temperature, which is the normal
temperature as measured by a mercury thermometer. The vertical axis on the right-side
of the sheet shows the absolute humidity of air (ω), also called the specific humidity,
which is the mass of water per kg of dry air. The curved lines give the relative humidity
(ϕ). The other properties given by psychrometric chart are:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 129

a. The welt-bulb temperature (Twb)


b. The enthalpy (h) in kJ/kg dry air
c. The specific volume (v) in m3/kg dry air

Depending on the initial temperature and humidity level of the air to be conditioned, the
desired condition may require heating, cooling, humidification, or dehumidification.
Figure 7.2 shows these four basic processes on the psychrometric chart. Air conditioning
processes usually involve combinations of the four basic processes such as the cooling
with dehumidification process shown in Figure 7.2.

Barometric pressure

Absolute humidity
Humidification

Cooling Heating

Cooling with Dehumidification


dehumidification

Dry bulb temperature

Figure 7.2. The psychromeric chart showing basic air-conditioning processes

Analyses of the different air-conditioning processes by using psychrometric charts can


be found in standard thermodynamics textbooks [1, 2]. The basic principles underlying
such analyses treat air as a mixture of two ideal gases; dry air and water vapour. It is
assumed here that the reader is familiar with these principles. The following sections
demonstrate the usefulness of the functions supported by Thermax for these analyses.

7.2. Cooling with dehumidification


The temperature and humidity level of air in hot and humid regions are higher than
normally required for human comfort. The needed cooling and dehumidification of the
air can be achieved by passing it over a humidifier followed by a cooling coil of a suitable
capacity. The air is first cooled to its dew-point temperature so that part of the airborne
water is condensed. The discharged air will be saturated with water vapour and at a lower
temperature. If the exit temperature of the dehumidified air is too low for human comfort,
or the humidity level is too high, the air is then heated to the required level of temperature
and humidity. The following example analyses the cooling-and-dehumidification process
so as to determine the rates of heat and moisture removal from the air. The example is
adopted from Cengel and Boles [1].
130 Mohamed M. El-Awad

Example 7.1. Cooling and dehumidification of air


Figure 7.3 shows a window air conditioner in which air enters at 1 atm, 30°C, and 80%
relative humidity at a rate of 10 m3/min. The part of the air moisture that condenses during
the process is removed at 14°C. Determine the rates of heat and moisture removal from
the air if it also leaves at 14°C as saturated air.

Cooling coil

2 1

Air

Condensate

Figure 7.3. Schematic diagram of a dehumidifier (adapted from Cengel and Boles [1])

The rate of heat removal ( Q ) is obtained from energy balance as follows:

Q  m a h1  h2   m wh f (7.1)

Where,  a is mass flow rate of dry air, h1 and h2 are values of enthalpies of humid air
m
at points 1 and 2, respectively, m w is the rate of moisture removal, and hf is the enthalpy
of saturated liquid water at T2. The rate of moisture removal is given by:

m w  m a 1  2  (7.2)

Where, and ω1 and ω2 are values of the absolute humidity at points 1 and 2, respectively.

 a ) involved in both Equations (7.1) and (7.2) is obtained


The mass flow rate of dry air ( m
by dividing the given volume flow rate by the specific volume of air at the specified
temperature and pressure, or:

V
a 
m (7.3)
va

Where, V is the volume flow rate of air and va is the specific volume of dry air which is
obtained from the ideal-gas law as follows:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 131

RT
va  (7.4)
Pa

Where, R is the gas constant for air (R=0.287 kJ/kg.K), T the absolute temperature and
Pa the partial pressure for dry air. The partial pressure Pa is given by:

Pa = P - Pv (7.5)

Equations (7.1) to (7.5) involve a number of thermodynamic properties of dry air, humid
air, or water vapour that have to be known in advance. Enthalpy of humid air (h) in
Equation (7.1) is calculated from:

h  c pa T  hv (7.6)

Where, cpa is the specific heat at constant pressure for dry air (cpa =1.005 kJ/kg.K) and T
is the temperature in oC. The absolute humidity (ω) in Equation (7.1) is determined from
the partial pressure of water-vapour in the air (Pv) according to:

0.622 Pv
 (7.7)
P  Pv

The partial pressure of water vapour (Pv) is determined from:

Psat
Pv  (7.8)
100

Where ϕ is the relative humidity and Psat is the saturation pressure of water at the given
temperature.

Excel model
The Excel sheet prepared for the solution of this example is shown in Figure 7.4. The
sheet reveals the formulae used in the calculations part. Figure 7.4 shows the calculated
values of the absolute humidity at points 1 and 2, which are calculated from Equation
(7.7) by using Thermax function PsySh_PDbRh. The values determined by this function
for the absolute humidity ω1 and ω2 are 0.02157 and 0.00995, respectively. Values of the
enthalpies h1 and h2, which are determined from Equation (7.9) by using the function
Psyh_PDbSh, are 85.266 kJ/kg and 39.216 kJ/kg, respectively. The value of saturated
liquid water hv is determined by using the function PsyhL_T* is 58.72 kJ/kg.
Substitutions in Equations (7.4) and (7.5) give the rate of water and heat removal as
0.1304 kg/min and 509.265 kJ/min, respectively.
* The function PsyhL_T returns the enthalpy of saturated liquid water at the given temperature in
o
C. This auxiliary function is not shown on Table 2.10.
132 Mohamed M. El-Awad

Figure 7.4. Excel sheet for cooling with dehumidification with Thermax

Figure 7.5 shows another Excel sheet that was prepared for the example by using the
“Thermotables” add-in developed at the University of Alabama [3]. Table 7.1 that
compares the values obtained by the two add-ins for different properties with their
corresponding values given by Cengel and Boles [1] shows good agreements between the
three solutions. The sheets can easily be modified to analyse the same process with
different input data.

Figure 7.5. Excel sheet for cooling with dehumidification with Thermotables

Table 7.1. Comparison of the present solution with that of Cengel and Boles [1]
Property Ref [1] Thermax Thermotables Deviation (%)
1 0.0216 0.0216 0.02165 -0.0
2 0.0100 0.0100 0.01000 -0.0
h1 85.4 85.266 85.4798 -0.157
h2 39.3 39.216 39.3489 -0.214
ma 11.25 11.225 11.2237 -0.221
hf 58.8 58.722 58.7936 -0.133
Q 511 509.265 510.076 -0.339
mw 0.131 0.130 0.1307 -0.479
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 133

7.3. Adiabatic mixing of two air streams


Mixing of unconditioned air with recirculated conditioned air is practiced in order to
reduce the consumption of energy in air-conditioning systems. Properties of the mixed
stream can be obtained by applying the principles of energy and mass conservation to the
mixing process (Figure 7.6). The following example that illustrates the use of Thermax
functions for the analysis of this process is also adopted from Cengel and Boles [1].

Example 7.2. Adiabatic mixing of two air streams


Saturated air leaving the cooling section of an air-conditioning system at 14oC at a rate
of 50 m3/min is mixed with the outside air at 32oC and 60% relative humidity at a rate of
20 m3/min. Assuming that the mixing process occurs adiabatically at a pressure of 1 atm,
determine the relative humidity, the dry-bulb temperature, and the volume flow rate of
the mixed stream.

Cold air 1

Mixing Mixed air


3
section

Outside air 2

Figure 7.6. Adiabatic mixing of two air streams (adapted from Cengel and Boles [1])

The analytical model


Applying conservation-of-mass to the mass flow rates of dry air in the three streams:

m a ,3  m a ,1  m a , 2 (6.9)

Similarly, for the mass of water-vapour:

3 m a,3  1m a,1   2 m a, 2 (6.10)

Elimination of m a ,3 by using Equation (6.9) yields the following relationship:

 a,11  m
m  a, 2 2
3 
m a,1  m a,2 
(6.11)

Conservation-of-energy gives:
134 Mohamed M. El-Awad

 a ,3 h3  m
m  a ,1h1  m
 a , 2 h2 (6.12)

Elimination of m a ,3 from Equation (6.12) leads to the following relationship:

m a,1h1  m a, 2 h2
h3 
m a,1  m a,2 
(6.13)

Excel model
Figure 7.7 shows the Excel sheet developed for this example. The data part stores values
of the temperatures, relative humidities, and flow rates of the two inlet streams as well as
value of the atmospheric pressure. Figure 7.7 reveals the formulae and the add-in
functions used in the calculations part.

Figure 7.7. Excel sheet developed for Example 7.2

The function PsySh_PDbRh is used to determine the specific humidities at states 1 and
2 from the known values of the pressure, dry-bulb temperature and relative humidity.
Similarly, the functions Psyh_PDbRh and Psyv_PDbRh are used to determine the
specific enthalpies and specific volumes at these states. The two inlet mass flow rates of
dry air, ma_1 and ma_2, are calculated from the respective volume flow rates and specific
volumes. Equations (6.9), (6.11) and (6.13) are then used to determine the mass flow rate,
specific humidity, and specific enthalpy at state 3. The dry-bulb temperature at state 3 is
determined by using the function PsyDb_PhSh while the relative humidity is determined
by using the function PsyRh_PDbSh. Table 7.2 compares the present results with those
provided by Cengel and Boles [1] who solved the example by using a psychrometric
chart. The figures on the table show minor deviations from the solution given by Cengel
and Boles [1].

7.4. The evaporative cooler


Evaporative cooling is an effective method for cooling the air in desert-like environments
where the air temperature is high and its humidity is low. When a hot and dry air is passed
through a bad soaked by water, the water vaporises into the air by taking the required
vaporisation energy from the air stream. This makes its temperature drops, while its
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 135

humidity increases, to more comfortable levels. The system requires only a fan to
circulate the air and a small pump to supply the water. Because of its low initial cost and
operating cost compared to the more common refrigerative cooler, the evaporative cooler
is preferred in many parts of the world particularly during the summer time. Note that the
lowest exit air temperature that can be reached by the evaporative cooler is the wet-bulb
temperature at the inlet condition. The following example is adopted from Moran and
Shapiro [2] to illustrate the use of Thermax functions for the analysis of evaporative
cooling.

Table 7.2. Verification of Example 6.2 results


Ref. [1] Thermax Deviation (%)
 a ,1
m 60.5 60.344 -0.258
 a,2
m 22.5 22.426 -0.327
h3 50.1 49.805 -0.588
T3 19 18.928 -0.379
3 89 88.759 -0.271
V3 70.1 70.001 -0.141

Example 7.3. Analysis of the evaporative cooler


Figure 7.8 shows a schematic diagram of a simple evaporative cooler. Air at 38oC and
10% relative humidity enters an evaporative cooler with a volumetric flow rate of 140
m3/min and exits the cooler as moist air at 21oC. Water is added to the soaked pad of the
cooler as a liquid at 21oC and evaporates fully into the moist air. There is no heat transfer
with the surroundings and the pressure is constant throughout at 1 atm. Determine (a) the
mass flow rate of the water to the soaked pad, in kg/h, and (b) the relative humidity of
the moist air at the exit to the evaporative cooler.
Water 21oC

1 2
o
T1 = 38 C o
ϕ1 = 10% T2 = 21 C

Soaked pad Boundary

Figure 7.8. Schematic of the evaporative cooler (Adapted from Moran and Shapiro [2])
136 Mohamed M. El-Awad

The analytical model


Neglecting energy transfer between the system and the surroundings and considering the
energy transfer between the air and water streams to take place within the control volume,
the energy equation simplifies to:

h2  h1   2  1 h f (7.14)

Where hf refers to specific enthalpy of the liquid stream entering the control volume. The
second term on the r.h.s. of Equation (7.14) is usually small compared to h1 and h2 and
can be neglected. Therefore, the enthalpy of the airstream remains nearly constant. Since
the constant wet-bulb temperature lines almost coincide with the constant-enthalpy lines,
the following approximation can also be made:
Twb,1 ≈ Twb,2 (7.15)

Given the exit temperature or relative humidity of the air leaving the air-cooler, state 2
can easily be determined. The amount of water m w required to cool air, initially at T1
and ϕ1, to a final temperature T2 can then be obtained by mass balance:

m w  m a  2  1  (7.16)

Excel model
Figure 7.9 shows the Excel sheet developed for this example with the input data and the
formulae used in the calculations part. The sheet uses Thermax property functions to
determine the enthalpy, specific humidity, specific volume, and wet-bulb temperature at
state 1 from the given values of the temperature and relative humidity. The specific and
relative humidities at state 2 are then determined from the given dry-bulb temperature
and the calculated wet-bulb temperature.

Figure 7.9. Excel sheet for Example 7.3

The mass flow rates of the dry air and the water are calculated based on Equations (7.3)
and (7.16), respectively. Table 7.3 compares the values obtained by Thermax functions
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 137

for the key parameters with their corresponding values given by Moran and Shapiro [2].
As the values in the table show, the differences between the two results are minor.

Table 7.3. Deviations of the present results from those given by Moran and Shapiro [2]
Ref [2] Thermax Deviation (%)
ω1 [kg vapour/kg dry air] 0.00408 0.00410 -0.490
va1 [m3/Kg dry air ] 0.887 0.890 -0.338
ma [Kg air/h] 2.63 2.623 0.266
ω2 [kg vapour/kg dry air] 0.011 0.011 0
mw [Kg water/h] 65.5 66.466 -1.475

7.5. Wet cooling towers


Steam-turbine power plants reject large quantities of heat from their condensers. The heat
is usually rejected into a nearby water reservoir or river, but if this is not possible it then
is rejected into the atmospheric air by using a cooling tower. Induced-draft cooling towers
are a type of cooling towers in which the warm cooling water is sprayed into a stream of
air induced into the tower by a large electric fan as shown in Figure 7.10. Part of the water
vaporises into the air and the required energy for its vaporization is drawn from the water
spray which reaches the tower bottom at a reduced temperature. Makeup water has to be
added to the cooling water in order to substitute the amount of water lost during the
vaporisation process. Both the required power for the fan and the amount of makeup
water depend on the temperature and humidity of the ambient air.

2 Humid air

3 Fan

Warm water

1
Dry
air

Cool water 4

Make-up
water
Figure 7.10. Schematic of the wet cooling tower (adapted from Cengel and Boles [1])
138 Mohamed M. El-Awad

Example 7.4. Cooling of a power plant by a cooling tower


The cooling water leaving the condenser of a power plant enters a wet cooling tower at
35°C at a rate of 100 kg/s. The water is cooled in the cooling tower to 22°C by air that
enters the tower at 1 atm, 20°C, and 60 percent relative humidity and leaves saturated at
30°C. Neglecting the power input to the fan, determine (a) the volume flow rate of air
into the cooling tower and (b) the mass flow rate of the required makeup water.

The analytical model


At steady operation, mass rates balances for the dry air and the water lead to:

m a,1  m a,2  m a (7.17)

m w,3  m a1  m w,4  m a2 (7.18)


Where m w refers to the mass flow rate of water. The makeup water is given by:

m Makeup  m w,3  m w, 4  m a  2  1  (7.19)

Neglecting heat transfer with the surroundings and the changes in kinetic and potential
energy, energy rates balance over the system leads to:

 m h   m h
In Out
(7.20)

m a h1  m w,3 hw,3  m a h2  m w,4 hw,4 (7.21)

Or,

 
m a h1  m w,3 hw,3  m a h2  m w,3  m Makeup hw, 4 (7.22)

Substituting for m Makeup from Equation (7.19) gives:

m w hw,3  hw, 4 
m a 
h 2  h1    2  1 hw, 4  (7.23)

Given the mass flow rate and temperature of the hot water and the desired temperature of
the cold water and the temperature and humidity of the air, the required mass flow rate
of the dry inlet air can be determined from Equation (7.23). The required volume for rate
of air can then be obtained from:

Va  m a / v a ,1 (7.24)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 139

Excel model
Figure 7.11 shows the Excel sheet developed for this example. The formulae used in the
analysis are shown in the figure with Thermax functions. The sheet uses three functions
from the Psy-group to determine the required properties of moist air, which are:
Psyh_PDbRh, PsySh_PDbRh, and Psyv_PDbRh. The sheet also uses one function
from the Wat-group, which is Wath_Tx, to determine the enthalpies of the warm and the
cold liquid water. Figure 7.11 reveals the formula that uses Equation (7.23) to calculate
the mass flow rate of air (ma). Based on this mass flow rate, the sheet calculates the
required volume flow rate of air (Va) to determine the required fan power and the required
flow rate of make-up water. This example is based on EXAMPLE 14–9 in Cengel and
Boles [1]. Table 7.4 that compares the values of the key parameters determined by
Thermax functions with their corresponding values shows a good agreement between the
two results.

Figure 7.11. Excel sheet for Example 7.4

Table 7.4. Verification of Thermax results for Example 7.4


Ref [1] Thermax Deviation (%)
ω2 kg H2O/kg dry air 0.0273 0.0272 -0.366
Air mass flow rate [kg/s] 96.900 97.580 0.702
v1 [m3/kg dry air] 0.842 0.844 0.238
Va [m3/s] 81.6 82.384 0.961
Makeup water [kg/s] 1.80 1.802 0.111

The Excel sheet developed for this example can be used to study the effect of the air
temperature and humidity on the required air flow rate and the amount of makeup water
as shown in Figure 7.12 and Figure 7.13. Figure 7.12, which was obtained by changing
the air temperature from 10oC to 30oC while keeping the air humidity constant at 60%,
shows that air flow rate increases from about 82 m3/s at the air temperature of 10oC to
nearly 175 m3/s at 30oC. The figure also shows that the required makeup water increases
140 Mohamed M. El-Awad

from about 1.6 kg/s to 2.2 kg/s in the same temperature range. Figure 7.13 that shows the
effect of the air humidity on the required air flow rate and make-up water was obtained
by changing the air-humidity from 10% to 60% while keeping the air temperature
constant at 20oC. For the given inlet and outlet temperature of the cooling water, the
figure indicates that the required air flow rate decreases to 62 m3/s, but the required
makeup water increases to about 1.9 kg/s if the air humidity drops to 10%.

200 2.4

2.2
Air volume flow rate (m3/s)

150

Make-up water (kg/s)


2

1.8
100
1.6

1.4
50 Air flowrate
1.2
Make-up water
0 1
0 10 20 30 40
Dry-bulb temperature (oC)

Figure 7.12. Effect of air dry-bulb temperature at ϕ = 60%


100 2

90 1.8
Make-up water (kg/s)
Air volume flow rate (m3/s)

80 1.6

70 1.4

60 Air-flow rate 1.2

make-up water
50 1
0 20 40 60 80
Relative humidity (%)

Figure 7.13. Effect of air humidity at T1 = 20oC


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 141

7.6. Design analysis of an air-conditioning system


Psychrometric analyses that use fluid property functions become more advantageous over
those with the psychrometric chart when dealing with parametric or optimisation analyses
of air-conditioning systems. This is demonstrated by the following design case which is
based on a design problem given by Moran and Shaprio [2].

The design case


The energy requirements of air-conditioning systems can be reduced by re-circulating
part of the discharged conditioned air and mixing it with the fresh unconditioned air.
Figure 7.14 shows a system for supplying a space with 2100 m3/min of conditioned air
at a dry-bulb temperature of 22oC and a relative humidity of 60% when the outside air is
at a dry-bulb temperature of 35oC and a relative humidity of 55%.

Recirculated air

Damper Damper 4
A B

4 4

1 5 6 3
Air-conditioned
Outside air space
2

Cooling Heating
coil coil
Condensate

Figure 7.14. Schematic diagram of the air-conditioning system (adapted from


Moran and Shapiro [2])

It is required to evaluate the effect of mixing part of the discharged air with the fresh
incoming air. Dampers A and B can be set to give three alternative operating modes:

(a) Both dampers closed (zero recirculated air).


(b) Damper A open and damper B closed. One-third of the conditioned air comes
from outside air.
(c) Both dampers open. One-third of the conditioned air comes from outside air.
One-third of the recirculated air bypasses the dehumidifier via open damper
B, and the rest flows through the damper A.

Which of the three operating modes gives the minimum cooling and heating loads?
142 Mohamed M. El-Awad

Option (a)
This option will be the reference case against which the savings of the other two modes
will be evaluated. Since there is no air re-circulation in this mode, both the cooling and
heating loads will be at their maximum values. The air-conditioning process consists of
only two operations; cooling and dehumidification process followed by a simple heating
process. Referring to Figure 7.14, the cooling load (Qc) and heating load (Qh) can be
calculated from:

Q c  m a1 h1  h2   m wh fg (7.25)

Q h  m a1 h3  h2  (7.26)
Where m a1 and h1 are the mass-flow rate of dry air and enthalpy of humid air at state 1,
h2 and h3 are the enthalpy of humid air at states 2 and 3, respectively, and m w is the rate
of water removal by condensation at state 2. The values of m a1 and m w are calculated
from:

V1 2100
m a1   (7.27)
va1 va1

m  a1 1  2 
w  m (7.28)

The condition of air at states 1 and 3 are known and its condition at state 2 is specified
by the fact that ϕ2 = 100% and ω2=ω3. Figure 7.15 shows the Excel sheet prepared for
this case.

Figure 7.15. Excel sheet for option (a)

The given data for T1, P1, ϕ1, ϕ2, T3, ϕ3, and Va3 is shown on the left side of the worksheet.
The specific humidities ω1 and ω3 are calculated by calling the add-in function
PsySh_PDbRh with relevant values of P, T and ϕ at the two points. At point 2, ω2 = ω3,
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 143

ϕ2=100% and P1=100 kPa. The temperature T2 is then determined by using the function
PsyDb_PRhSh. Values of the enthalpy h1, h2, and h3 are calculated by calling the
function Psyh_DbSh. The auxiliary function PsyhL_T is also used here (cell I10) to
determine the temperature of saturated liquid water at T2. Substituting in Equations (7.25)
and (7.26), the values obtained for the cooling and heating loads (Qc =Qout12 and Qh =
Qin23) determined by the Excel sheet for this option are 112,696.7 kJ/min and 20,227.4
kJ/min, respectively.

Option (b)
In this case, two-thirds of the discharged air is recirculated through damper A and only
one-third of the conditioned air comes from outside air. As shown later, this mode reduces
the cooling load, but keeps the heating load unchanged. Referring to Figure 7.14, the
mass flow rate of dry air at state 3 is given by:
V3 2100
m a 3   (7.29)
va3 va3

The mass flow rates of dry air at states A, 1, 2, and 5 are given by:

 aA  m
m  a3 / 3 (7.30)

 a1  m
m  a3 / 3 (7.31)

 a5  m
m  a2  m
 a3 (7.32)

The condition of air at the states 1, 2, and 3 are the same as in case (a). The condition at
state 4 is specified by the given values of T = 27oC and ϕ=49%. Assuming adiabatic
mixing, the condition of the air after the mixing process at state 5 is determined as
follows:

h5  m  A h4  / m
 1h1  m 5 (7.33)

5  m  A4  / m
 11  m 5 (7.34)

The rate of water removal in the dehumidifier is calculated from:

m  a3 5  2 
w  m (7.35)

The cooling load and heating loads for this case are calculated from:

Q c  m a3 h1  h2   m wh fg (7.36)
144 Mohamed M. El-Awad

Q h  m a3 h3  h2  (7.37)

Figure 7.16 shows the Excel sheet prepared for this case. In addition to the input-data of
the sheet shown in Figure 7.15, the input-data part of this sheet provides the condition at
point 4. The mass flow rate of outside air (mflowa_1) is reduced to one third of the total
flow rate and two-thirds of the total the mass flow rate passes through damper A
(mflowa_A). The calculation part is also expanded to include the calculations of humidity
and enthalpy of the air mixture at point 5. The cooling and heating loads are now
determined according to Equations (7.36) and (7.37). As Figure 7.16 shows, the cooling
and heating capacities for this option are 63,575.9 kJ/min and 20,227.4 kJ/min,
respectively. As expected, this option reduced the cooling capacity compared to option
(a), but the heating capacity remained unchanged.

Figure 7.16. Excel sheet for option (b)

Option (c)
In this option, one-third of the conditioned air bypasses the dehumidifier via the open
damper B. Another one-third of the air-flow is recirculated via damper A while the
remaining air comes from the outside. This mode will reduce the cooling as well as the
heating capacities of the system. Referring to Figure 7.14, the mass flow rate of dry air
at state 3 is:

V3 2100
m a 3   (7.38)
va3 va3

The mass flow rates of dry air at states A, B, and 1, which are equal are, given by:

 aA  m
m  aB  m
 a1  m
 a3 / 3 (7.39)

Conservation of mass applied over the mixing processes after damper A and damper B
gives:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 145

 a5  2m
m  a3 / 3 (7.40)

 a2  m
m  a5 (7.41)

 a6  m
m  a3 (7.42)

The air conditions at the states 1, 3 and 4 are the same as in case (a) and case (b). The
condition at state 2 is also specified by the fact that ϕ2 = 100% and ω2=ω3. Assuming
adiabatic mixing, the condition after the mixing process at state 5 is determined by:

h5  m  A h4  / m
 1h1  m 5 (7.43)
5  m  A4  / m
 11  m 5 (7.44)

Similarly, the condition at state 6 is determined as follows:

h6  m  B h4  / m
 2 h2  m  a6 (7.45)

6  m  B4  / m
 22  m 6 (7.46)

The rate of water removal in the dehumidifier is calculated from:

m  a5 5  2 
w  m (7.47)

Finally, the cooling and heating loads for this case are calculated from:

Q c  m a 5 h1  h2   m w h fg (7.48)

Q h  m a3 h3  h6  (7.49)

Figure 7.17 shows the Excel sheet prepared for this case. Compared to the sheet shown
in Figure 7.16, the calculation part is expanded to calculate the mass flow rates after
damper B and at states 5 and 2. The additional calculations also include enthalpy and
humidity values and the temperature at state 6. The rates of cooling and heating capacities
calculated for this option are 50,570.1 kJ/min and 7,173.1 kJ/min, respectively. These
figures indicate that both the cooling and heating capacities of this option are significantly
less than those of options (a) and (b). Table 7.5 summarises the results of the cooling and
heating capacities for the three options. Compared to option (a), options (b) and (c) save
about 37% and 57% of the total load, respectively. Therefore, option (c) is the preferred
mode of operation.
146 Mohamed M. El-Awad

Figure 7.17. Excel worksheet for option (c)

Table 7.5. Cooling and heating capacities for the three options
Option Cooling capacity Heating capacity Total load Saving [%]
[kJ/min] [kJ/min] [kJ/min]
(a) 112,696.7 20,227.4 132,924.10 -
(b) 63,575.2 20,227.4 83,802.60 36.95
(c) 50,570.1 7,7173.1 57,743.20 56.56

Table 7.6 lists Thermax function used in the analysis of option (c) of the design case;
which is the case that involves the largest number of state points. As the table shows,
from the total of 18 functions provided by the “Psy” group of Thermax, only six functions
have been used in this design analysis. This indicates the adequacy of the add-in functions
in this group for a wide range of air-conditioning and psychrometric analyses.

Table 7.6. Thermax function used in the analysis of design case (c)
Function Usage
PsySh_PDbRh 1 ,  3 ,  4
PsyDb_PRhSh T2
Psyh_PDbSh h1 , h2 , h3 , h4
PsyDb_PhSh T5, T6
PsyPsat_T Pv3
PsyhL_T Qc

7.7. Closure
This chapter demonstrated the usefulness of Excel with Thermax property functions for
psychrometric analyses. The results obtained with Thermax functions for the analyses of
selected air-conditioning processes and systems confirmed the convenience and accuracy
of these functions. The chapter also considered a design case that showed how the
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 147

functions can be used for design optimisation analysis of an air-conditioning system.


Unlike the psychrometric charts which are limited to a single pressure, usually the
atmospheric pressure, the property functions can be used at different pressure levels
provided that the underlying ideal-gas assumption applies. The analyses presented in this
chapter can also be performed by using the “Thermotables” add-in developed at the
University of Alabama [3].

References
[1] Y.A. Cengel, and M.A. Boles, Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007
[2] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley, & Sons. Inc. 2006.
[3] The University of Alabama, Website of Excel in Mechanical Engineering,
https://www.me.ua.edu/ExcelinME/
148 Mohamed M. El-Awad
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 149

Chapter 8

Development of a general Excel sheet


for combustion analyses
150 Mohamed M. El-Awad

Like the analyses of non-reacting systems considered in the previous chapters,


combustion processes can be analysed on the basis of mass and energy conservation
principles. However, combustion analyses require the thermodynamic properties of
multiple reactants (fuel, air, and water-vapour) and multiple combustion products (CO2,
CO, H2O, O2, and N2). Determining these properties at different temperatures makes
combustion analyses tedious for hand calculations using property tables. This chapter
shows how the property functions provided by Thermax can be used to develop a general
Excel sheet for combustion analyses and how the sheet can be used to determine the air-
fuel ratio, the dew-point temperature of the combustion products, and the amount of heat
released from combustion of conventional and alternative fuels. Excel’s Goal Seek
command makes the sheet particularly useful for performing combustion analyses that
require iterative solution. In this respect, the chapter illustrates the use of Goal Seek for
determining the fuel’s adiabatic flame temperature (Taf) for steady-flow combustion.

8.1. The basic equations for combustion analyses


This section develops the basic equations for the combustion of a hydrocarbon fuel in air.
The equations can be used to determine: (i) the amount of air needed for the combustion
process (or the air-fuel ratio), (ii) the temperature at which condensation of water vapour
in the combustion products occurs (the dew-point temperature of combustion products),
(iii) the fraction of condensed water-vapour at temperatures lower than the dew-point
temperature, and (iv) the amount of heat released by the combustion process.

8.1.1. The combustion equation for a hydrocarbon fuel in air


Ideal combustion of a hydrocarbon fuel CαHβ in pure oxygen is expressed by the
following equation:

C H   aO2  bCO2  cH 2O (8.1)

Balances of the number of C moles, O2 moles, and H2 moles on both sides of the equation
lead to:

b  (8.2)

c /2 (8.3)

a bc/2 (8.4)

Substituting from Equations (8.2) and (8.3) in Equation (8.4) leads to:

a    / 4 (8.5)

Therefore, the chemical equation for combustion in pure oxygen can be written as:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 151

 
C H   (  )O2  CO2  H 2O (8.6)
4 2

In actual combustion processes oxygen comes from atmospheric air which, in addition to
oxygen, contains nitrogen and traces of other gasses as well as water vapour. In
combustion analyses, air is assumed to contain 3.76 moles of nitrogen for every mole of
oxygen. Since nitrogen is an inert gas, the same number of N2 moles that enter the
combustion process appear in the products side of the equations, i.e.:

  
C H   (  )O2  3.76 N 2   CO 2  H 2 O  3.76 (  )N2 (8.7)
4 2 4

Equation (8.7) expresses the mass balance in stoichiometric combustion, i.e. combustion
with the exact amount of air needed for complete reaction. Actual combustion processes
can take place in excess air or in excess fuel. When excess air is provided, the amount of
excess oxygen and excess nitrogen will appear in the products side of the equation. The
equation for combustion of a hydrocarbon fuel in excess amount of air can be written as:

 
C H    (  )O2  3.76 N 2   CO 2  H 2 O  dN 2  eO2 (8.8)
4 2

Where, γ indicates the amount of excess air provided for combustion:

Number of moles of actual air provided for combustion


 (8.9)
Number of moles of air needed for stoichiometric combustion

According to the value of γ, the air-fuel mixture is called:

γ = 1 stoichiometric mixture
γ > 1 lean mixture
γ < 1 rich mixture

Given the values of α, β, and γ, the two unknowns (d, e) can be determined from atom
balance of oxygen and nitrogen. From the balance of nitrogen atoms:


d  3.76 (  ) (8.10)
4

From oxygen balance:

 
 (  )   e (8.11)
4 4
152 Mohamed M. El-Awad

Rearranging the equation,

 
e  (  1)    (8.12)
 4

Substituting for d and e in Equation (8.8), the combustion equation becomes:


C H    (  )O2  3.76 N 2  
4
    
CO2  H 2O  3.76    N 2    1  O2 (8.13)
2  4  4

Equation (8.13) assumes the combustion air to be dry, but it can be modified to account
for the effect of air-borne moisture. Like nitrogen, air humidity increases the number of
moles in both the reactants and products sides of the chemical reaction equation. Taking
the number of moles due to air moisture alone as Nv,air, the combustion equation becomes:


C H    (  )O2  3.76 N 2   N v , air H 2O 
4
   
CO2  H 2O   (  )3.76N 2  (  1)   O2  N v , air H 2O (8.14)
2 4  4

The mass ratio of air to fuel present in a fuel-air mixture is referred to as the air-fuel ratio
(AFR). From Equation (8.14), the AFR for a hydrocarbon fuel is given by:


(  )M O 2  3.76M N 2   N v, air M H 2O
AFR  4 (8.15)
M fuel

Where, MO2, MN2, MH2O, and Mfuel are, respectively, the molar masses of oxygen, nitrogen,
water vapour, and the fuel. If exactly enough air is provided to completely burn all of the
fuel (i.e., γ = 1), the air-fuel ratio is known as the stoichiometric air-fuel ratio (AFRs).

The combustion equation can also account for the situation when there is excess fuel (i.e.
a rich mixture) or when combustion is incomplete. Combustion equations for other types
of fuels, such as coal or alcohols that include sulphur, nitrogen, or oxygen in their
molecules, can be developed following the same procedure. For these aspects, the reader
can refer to the more complete treatments given in standard textbooks [1-3].

8.1.2. The dew-point temperature of combustion products


The dew-point temperature of the combustion products is the temperature at which water-
vapour in the products begins to condense, which occurs when the partial pressure of the
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 153

water-vapour in the products equals the corresponding saturation pressure. The partial
pressure (Pv) of the water vapour in the products is calculated from:

Nv
Pv  Ptotal (8.16)
Np

Where, Nv and Np are the number of moles of water vapour in the products and the total
number of moles of products, respectively, and Ptotal is the total pressure. Two cases can
be considered in this respect, which are: (i) combustion with dry air and (ii) combustion
with moist air.

a) Dew-point temperature for products of combustion with dry air


From dry air, Equation (8.13) yields:


Nv  (8.17)
2

   
Np     (  )3.76  (  1)    (8.18)
2 4  4

Once the values of Nv and Np are determined for a given hydrocarbon fuel and the amount
of excess air, the partial pressure (Pv) of water-vapour in the combustion products can be
calculated from Equation (8.16). The dew-point temperature (Tdp) is then the
corresponding saturation temperature:

Tdp  Tsat ( Pv ) (8.19)

The dew-point temperature Tdp can be determined from the property tables for saturated
water at the calculated partial pressure Pv. The function “WatTs_P(P)” provided by
Thermax determines the dew-point temperature at a given pressure.

b) Dew-point temperature for products of combustion with humid air


Equations (8.16) and (8.19) can still also used to obtain the dew-point temperature for
the products of combustion with humid air, but to allow for atmospheric moisture, Nv and
Nproducts should now be determined from:


Nv   N v , air (8.20)
2

    
N p      N v , air    (  )3.76  (  1)    (8.21)
2  4  4

The value of Nv,air itself is given by:


154 Mohamed M. El-Awad

M air
N v , air  N air (8.22)
Mv

Where, Nair is the number of moles of oxygen and hydrogen in the combustion air, ω is
the specific or absolute air humidity, and Mair and Mv are the molar mass for air and water
vapour, respectively. The absolute humidity (ω) can be calculated from the air
temperature (Tair) and pressure (Pair) and relative humidity (ϕ). First, the partial pressure
of air-borne water vapour (Pv,air) and absolute humidity (ω) are calculated from:

Pv , air  Pv, sat (Tair ) (8.23)

mv, air 0.622 Pv, air


  (8.24)
mair Pair

Therefore, taking the molar mass of air and water vapour as 29 and 18.02, respectively,
Equation (8.22) reads:

29
N v , air  N air  1.611N air (8.25)
18.02

8.1.3. Condensation of water vapour in the combustion products


If the products temperature falls below the dew-point temperature of the products (Tdp),
some of the water will be condensed. The faction of condensed water, which depends on
the products temperature (Tex), must be taken into consideration in energy balance and
second-law analyses of the combustion process. Suppose that the amount of water vapour
present in the combustion products (including air humidity) is Nv. Then the amount of
water vapour (Nwv) can be determined as follow:

N wv  N v Tex >= Tdp (8.26)

N wv 

Pr N p  N v  Tex < Tdp (8.27)
1  Pr

Where Pr is the ratio of the saturation pressure at the products’ temperature to the total
pressure:

Pr  Psat @ Tex / Pp (8.28)

The amount of condensed water (Nwl) is then given by:

N wl  N v  N wv (8.29)
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 155

8.1.4. The first-law of thermodynamics for steady-flow combustion


Combustion is an exothermic chemical reaction during which the fuel’s oxidation
releases a large quantity of thermal energy. The combustion process usually takes place
in the cylinder of an internal-combustion engine (a closed system) or in the combustion
chamber of a gas-turbine or a jet engine (steady-flow devices). The amount of heat and
work exchange between the combustion device and the surroundings can be determined
by applying the first-law of thermodynamics to the combustion process. The following
treatment deals with steady-flow combustion.

Neglecting kinetic and potential-energy changes, the first-law of thermodynamics for a


steady-flow combustion process states [1]:

 n  n
~ ~
Q  W  P hP  R hR (8.30)
P R

Where Q and W are the rate of heat and work transfer, respectively, n P and n R are the
~
molal flow rates of the product P and the reactant R, respectively, and h is the molal
specific enthalpy of the reactants or products component. The sign convention adopted
in Equation (8.30) is that heat out of the system and work into of the system are negative.

For combustion analyses, the equation is more conveniently expressed per mole of fuel
by dividing each term of the equation above by the molal flow rate of the fuel. Thus, the
following equation is obtained:

N N
~ ~
Q W  P hP  R hR (8.31)
P R

Where NP and NP are the number of moles of each component in the reactants and
products of combustion, respectively. Since values of enthalpies appear in Equation
(8.31), and not enthalpy differences, combustion analyses require a common reference
~
for enthalpy values for all reactants and products. Accordingly, the molal enthalpy h is
calculated from:
~ ~ ~
h  h f0  h
(8.32)

0~
Where h f is the formation enthalpy of the component at standard condition of 25oC and
~
1 atm and h is the deviation of its specific enthalpy at the relevant state from that at the
standard condition. Substitution from Equation (8.32) in Equation (8.31) gives:

 N h    N h~ 
~ ~ ~
Q W  P
0
f  h P R
0
f  h R
(8.33)
P R
156 Mohamed M. El-Awad

~ ~
Thermax provides property functions that determine the values of both h f0 and h for
different fuels and products of combustion (refer to Table A.4 in Appendix A).

8.2. Development of the general sheet for combustion analyses


Thermax enables a general Excel sheet for combustion analyses to be developed by
applying the equations presented in the previous section. Given the name and condition
of the hydrocarbon fuel and the amount and condition of combustion air, the sheet
calculates the air-fuel ratio, dew-point temperature of the combustion products, amount
of water vapour in the products, and amount of heat-transfer from combustion. The
following example shows how the sheet is developed using Thermax functions.

Example 8.1. Combustion of octane in excess humid air


Octane gas (C8H18) at 298K is burned steadily with 20 percent excess air at 330K, 1 atm,
and 55% relative humidity. Assuming combustion is complete and the products leave the
combustion chamber at 600K, determine (i) the air-fuel ratio, (ii) the dew-point
temperature of the combustion products and (iii) the heat transfer for this process per unit
mass of octane.

Solution
This example represents a general case that involves most of the factors involved in
energy analyses of combustion processes. Therefore, the Excel sheet prepared for solving
this example needs minor modifications to deal with other fuels or other inlet conditions
of the fuel and combustion air. Figure 8.1 shows the Excel sheet for combustion analyses
and Figure 8.2 reveals Excel formulae and Thermax functions used in the sheet.

Figure 8.1. The general sheet for combustion analyses (1st law analysis)

Figure 8.1 shows that the sheet has three main parts:

a) the problem data (columns B to F),


b) the common combustion calculations using Thermax functions (columns H
to L), and
c) the final results (columns N to R).
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 157

(a)

(b)

(c)
Figure 8.2. Excel formulae and Thermax functions used in the general sheet for
combustion analyses

(a) The data part


As shown in Figures 8.1 and 8.2.a, the problem data occupies the left side of the sheet
and specifies the type of fuel (Fuel) and its inlet temperature (Tin_fuel) and the condition
of combustion air, i.e. amount of excess air (γ), inlet-air humidity (Φ), inlet temperature
(Tin_air), and inlet pressure (Pin_air). The data part also shows the exit temperature of
the combustion products (T_ex).
158 Mohamed M. El-Awad

(b) The calculations part


The calculations part uses Thermax functions to determine the fuel’s values of α and β,
molar mass (M), and enthalpy of formation (hf_fuel) as shown in Figure 8.2.a. The
enthalpy of formation for each component in the reactors and the products are determined
by calling the function Chm_f2(“fuel”, “hf0”). Based on the fuel type (Fuel), the amount
of excess air (γ), and air-humidity (Φ), the sheet calculates the number of moles for each
component of the reactors (cells I2 to I5) and the products of combustion (cells I8 to I12).
In this respect, it should be noted that complete combustion is assumed, i.e. the number
of CO2 moles in the products in cell I8 (CO2_P) is made equal to α as required by
Equation (8.2). Accordingly, the number of moles of CO (CO_P) is zero. When dealing
with incomplete combustions this part has to be modified manually.

The deviations in molar enthalpies of the reactants at their inlet temperatures from those
at the standard temperature of 298K (ΔhR_fuel, ΔhR_O2, ΔhR_N2) are determined by
calling the Chm_f3(“Gas”, “hm”, TK) function in cells L2 to L5 as shown in Figure
8.2.b. Since in this case both the fuel and air enter the combustion process at 298K, their
enthalpy changes from the reference temperature in cells are all determined as zeros. This
is not the case with the changes in molar enthalpies of the products at the discharge
temperature from those at the standard temperature (ΔhP_CO2, ΔhP_H2O, ΔhP_N2,
ΔhP_O2, ΔhP_CO), which are determined in cells L8 to L12.

(c) The results part


The results part is shown on the right side of the sheet. As shown in Figure 8.2.b, the air
-fuel ratio is determined from Equation (8.15) as 18.23. The partial pressure for the water
vapour in the combustion air (Pv_R) and its specific humidity (ω) are then calculated and
used in Equation (8.19) to determine the dew-point temperature of the combustion
products (Tdp), which in this case is 51.8oC. Equation (8.27) is used to determine the
amount of condensed water vapour in the products (Nwl_ex), which is zero in this case.
Finally, the amount of heat transfer from the combustion process (Q) is determined from
Equation (8.33) in cell R8 and the relevant Thermax functions are shown in Figure 8.2.c.
Figure 8.1 shows that the sheet determines Q as -4,377,603.508 kJ/kmol of the fuel.

8.3. Using the general sheet for combustion analyses of various fuels
The combustion sheet developed in the previous section is general and requires only the
problem data part to be modified for complete combustion in stoichiometric or excess air
of all the fuels listed in Table A.4. For these fuels, the sheet allows the effects of the
following factors on the combustion process to be analysed:

a) The inlet temperature and pressure of the fuel and combustion air
b) The amount of excess air provided
c) The humidity of combustion air
d) The discharge temperature of the combustion products
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 159

The following examples illustrate the usefulness of the sheet for analysing various
scenarios of steady-flow combustion.

Example 8.2. Heat release in stoichiometric combustion of propane


A spark-ignition engine operating on stoichiometric propane fuel (C3H8) burns 0.00005
kg of fuel in each cylinder during each cycle. When combustion starts at the end of
compression, the temperature and pressure in the cylinder are T = 700 K and P = 2000
kPa, respectively. Exhaust temperature leaving the cylinder after combustion is Tex=1200
K. Calculate (i) the air-fuel ratio and (ii) the combustion heat input into each cylinder
during each cycle.

This example is adopted from Example 4.4 in Pulkrabek [4].

Solution
Figure 8.3 shows the general sheet with the modifications required by this example that
include the type of fuel (Fuel=C3H8) and inlet temperature (Tin_fuel=700K), the amount
of excess air (γ=0%), air inlet temperature and pressure (Tin_air =700K, Pin_air =101
kPa), and the exit temperature of the combustion products (T_ex = 1200 K).

Figure 8.3. Excel sheet for stoichiometric combustion of propane

The air-fuel ratio (AFR) is determined as 15.57 and the amount of heat released (Q) as
1,577,317 kJ/kmol of fuel. The combustion heat input per 0.00005 kg of the fuel burnt in
each cylinder (Qcyl) during each cycle is calculated from:

0.00005
Qcyl  Q kJ
M

Where M is the molar mass of the fuel, which is determined by Thermax function
Chm_f2(“fuel”, “M”). Note that the formula that determines Qcyl is shown in the formula
window in Figure 8.3. The calculated value of Qcyl is 1.79 kJ, while that given by
Pulkrabek [4] is 1.77 kJ.
160 Mohamed M. El-Awad

Example 8.3. Heat transfer from the combustion of vapour and liquid ethanol
One alternative to using petroleum or natural gas as fuels is ethanol (C2H5OH), which is
commonly produced from grain by fermentation. Consider a combustion process in
which ethanol is burned with 120% theoretical air in a steady flow process. The reactants
enter the combustion chamber at 25°C, and the products exit at 60°C, 100 kPa. Calculate
the heat transfer per kilomole of ethanol for: (a) vapour ethanol and (b) liquid ethanol.

This example is based on Problem 14.45 given by Sonntag, Borgnakke and van Wylen
[3].

Solution
(a) Figure 8.4 shows the modified sheet for the case of vapour ethanol in which the fuel
name has been entered as “C2H5OH”, which is Thermax name for ethanol. The inlet
conditions for the fuel and combustion air have also been adjusted according to those
given in the problem. Both the fuel and air enter the combustion process at 298K and,
therefore, their enthalpy changes from the reference temperature calculated in cells L2 to
L5 are all zero. As Figure 8.4 shows, the value determined by the sheet for the heat
transfer (Q) is 1,254,171 kJ/kmol of fuel. The value given by Sonntag, Borgnakke and
van Wylen [3] for Q is 1,256,904 kJ/kmol fuel.

Figure 8.4. Heat transfer from combustion of vapour ethanol

(b) Since liquid ethanol is included as a separate substance (C2H5OHL) in list of


reactants supported by Thermax shown in Table A.3, the sheet developed for part (a) can
be used for liquid ethanol by changing the fuel’s name from “C2H5OH” to “C2H5OHL”
as shown in Figure 8.5. For the gaseous reactants, O2, N2 and H2O, the relevant values
of enthalpy changes in cells L3 to L5 are automatically adjusted by the sheet, but for the
liquid fuel the enthalpy change (ΔhR_fuel) in cell L2 has to be set to zero by hand. In this
case, the value for the heat transfer determined by the sheet is 1,211,791 kJ/kmol of fuel
while the value given by Sonntag, Borgnakke and van Wylen [3] is 1,214,524 kJ/kmol
fuel.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 161

Figure 8.5. Heat transfer from combustion of liquid ethanol

Example 8.4. Combustion of methane gas with preheated air


Figure 8.6 provides data for a boiler and air preheater operating at steady state. Methane
(CH4) enters the boiler at 25oC, 1 atm is burned completely with 170% of theoretical air.
Combustion gases from the boiler enter the air preheater at 797oC and exit at 662 oC.
Ignoring stray heat transfer and kinetic and potential energy effects, determine the
temperature, in oC, of the combustion air entering the boiler from the preheater. If the
steam pressure is 8 MPa, what is the mass of steam produced per one kg of the fuel?

CH4

T=?
Air
Boiler
Preheater

Combustion 797oC
gas

Feedwater Steam

Figure 8.6. Schematic for Example 8.4 (adapted from Moran and Shapiro [2])

This example is based on Problem 13.38 in Moran and Shapiro [2].

Solution
In this example, the temperature of the combustion air is not known but has to be
determined from energy balance over the preheater. Applied over the preheater, the first
law dictates that the energy gained by the fresh combustion air (oxygen plus nitrogen) is
balanced by the heat lost by the combustion products, i.e.:

 ~  ~

  nh  
  nh   ~ ~
 N O 2 hO 2  N N 2 hN 2   N
T
~
O 2 hO 2
~
 N N 2 hN 2 
25 o C (8.34)
 P 797 o C  P 662 o C
162 Mohamed M. El-Awad

The left side of Equation (8.34) involves a known temperature and, therefore, can be
determined in a straight forward manner from the products enthalpies. However, the right
side involves the air temperature after the preheater, T, which is unknown. Therefore,
Equation (8.34) has to be solved by iteration to determine the temperature of preheated
air (T). The solution can be assuming an initial value for T and then use Goal Seek to find
the value that makes the difference between the two sides of Equation (8.34) equal to
zero by iteration.

Figure 8.7 shows the general sheet which has been modified according to the type of fuel
(CH4), fuel inlet temperature (Tin_fuel = 298K), amount of excess air (γ = 70%) and
temperature of the products after combustion (T_ex = 1070K). The temperature of
combustion air after the preheater (Tin_air) is initially assumed to be 700K. The middle
part of the sheet that involves the heat-release calculations remains unchanged.

Figure 8.7. Heat release for combustion of methane with preheated air

According to the assumed inlet-air temperature, the amount of heat released (Q) is
564,044 kJ/kmol fuel. The extension of the sheet in the results part (cells R6 to R16)
involves the calculations for energy balance over the air preheaters according to Equation
(8.34). The molar enthalpies of the products of combustion leaving the preheater are
determined at 935K (797oC) and the molar enthalpies of oxygen and nitrogen in the
combustion air are determined at the assumed temperature of T_ex=700K.

Figure 8.7 shows that the assumed value of T_ex leads to a difference (Diff) between the
energy lost by the combustion products (Δh_products) and the energy gained by the air
(Δh_air) of 114,908.5 kJ. The correct temperature of the preheated air at which the value
of Diff=0 can be found by Goal Seek. Figure 8.7 also shows the required set-up for Goal
Seek. The solution determined by Goal Seek is shown in Figure 8.8. Figure 8.8 shows
that the value of Tin_air that satisfies energy balance is 467.932K or 194.93oC. The value
obtained by Moran and Shapiro [2], who used the Interactive Thermodynamics (IT)
software to perform the iterative solution, is 196.4oC.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 163

Figure 8.8. Goal Seek solution for the temperature of preheated air

The mass of steam (msteam) produced by burning one kg of fuel is given by:

Q
msteam  kg/kmol fuel (8.35)
M h fg

Where, M is molar mass for methane and hfg is the latent heat of vaporisation for water
at the boiler pressure (8 MPa), which can be determined by using Thermax functions in
the group for saturated water. The sheet first determines the saturation temperature at 8
MPa by calling the function “WatTsat_P” from which hfg is determined by calling the
function “Wathfg_T”. The formula window in Figure 8.8 reveals the formula in cell R16
that applies Equation (8.35) to calculate the mass of steam produced from the combustion
heat per 1 kg of methane, which is 19.4 kg.

8.4. Determining the adiabatic flame temperature with Goal-Seek


The adiabatic flame temperature (Taf), or adiabatic combustion temperature, is the
temperature reached in the limiting case when there is no heat transfer or work interaction
with the surroundings. In this case, Equation (8.31) leads to:

N N
~ ~
P hP  R hR
P R
(8.36)

Although the right side in above equation can be evaluated directly from the initial state
of the reactants, the temperature at which the enthalpy of the products equals that of the
reactants, which is Taf, cannot be determined directly because the individual products
have different enthalpies at the same temperature. Therefore, an iterative process is
required in order to determine Taf. The following examples demonstrate the procedure of
determining Taf by using Goal Seek with the general combustion sheet.
164 Mohamed M. El-Awad

Example 8.5. The adiabatic flame temperature for octane


Find the adiabatic flame temperature of octane (C8H18) burned in a small turbocharged
automobile engine with 20% excess air in dry air. It can be assumed that the reactants are
at a temperature of 427oC (700 K) after the compression stroke.

This example is based on Example 4.5 in Pulkrabek [4].

Solution
Figure 8.9 shows the sheet developed for this example by modifying the data part of the
general sheet to suit this example. Figure 8.9 shows the heat transfer at an assumed
exhaust temperature (T_ex) of 1200 K, which is -3,768,068 kJ per kmol of fuel. We can
now use Goal Seek to determine the value of T_ex at which Q becomes zero. Figure 8.9
also shows Goal Seek set up for this task and Figure 8.10 shows the solution found by
Goal Seek. The solution, which is 2475.27 K, is slightly higher than that given by
Pulkrabek [4], which is 2419K.

Figure 8.9. Determining the adiabatic temperature of octane

Figure 8.10. Goal Seek solution for the adiabatic flame temperature of octane

8.5. Dealing with unsupported fuels


All the previous examples considered complete combustion of fuels that are supported
by the Thermax add-in. Two examples are given is section to show how the general
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 165

combustion sheet can be modified so as to deal with fuels not supported by the add-in
and with incomplete combustion.

Example 8.6. The heat-transfer from stoichiometric combustion of hexane


Determine the heat transferred from steady-flow combustion of hexane (C6H14) in dry
stoichiometric air when both the fuel and air enter the combustion chamber at 25 oC and
101 kPa and products of combustion cooled to 25oC.

Solution
Figure 8.11 shows the Excel sheet prepared for this example. Note that the amount of
excess air (γ) and air humidity (Φ) have been assigned a zero value and all the
temperatures have been assigned a value of 298K. Since hexane is not included in the list
of fuels supported by Thermax functions, its values of carbon and hydrogen contents (α
and β), molar mass (M), and enthalpy of formation (hf_fuel) have to be entered manually.
For hexane, the molar mass is 86.17 and enthalpy of formation for hexane is -4163.2
kJ/kmol. Figure 8.11 shows that 4,298,234 kJ of heat are released by the combustion of
one kmol of hexane.

Figure 8.11. Excel sheet for determining heat-transfer from stoichiometric combustion
of hexane

Example 8.6. The heat-transfer from combustion of liquid propane


Liquid propane (C3H8) enters a combustion chamber at 25°C at a rate of 0.05 kg/min
where it is mixed and burned with 50 per cent excess air that enters the combustion
chamber at 7°C. An analysis of the combustion gases reveals that all the hydrogen in the
fuel burns to H2O but only 90 per cent of the carbon burns to CO2, with the remaining 10
per cent forming CO. If the exit temperature of the combustion gases is 1500 K, determine
the rate of heat transfer from the combustion chamber.

Solution
Apart involving a fuel that is not supported by the add-in, this example, which is adopted
from Example 15.6 in Cengel and Boles [1], also differs from the previous examples in
that combustion is incomplete. Since gaseous propane is supported by the add-in
functions, the fuels carbon content (α), hydrogen content (β) and molar mass M) can be
166 Mohamed M. El-Awad

obtained from those of gaseous propane. The enthalpy of formation for liquid propane
h~  can also be obtained from that of gaseous propane h~ 
0
f l
0
f g by using the following
relationship:

h~   h~ 
0
f l
0
f g  Mh fg

Where, M and hfg are, respectively, the molar mass and heat of vaporisation for propane.
Thermax provides functions for both M and hfg of the supported fuels. Figure 8.12 shows
the modified sheet in which the name of the fuel has been entered as C3H8, but the
formula in cell F8 has been modified to apply the above equation as shown in the formula
window.

Figure 8.12. Excel sheet for heat release in combustion of liquid propane

To deal with incomplete combustion, the formulae that determine the number of CO2 and
CO moles in the products, i.e. cells I8 and I9, have to be overwritten by the values 2.7
and 0.3, respectively. Finally, the rate of heat transfer from the combustion chamber per
0.05 kg/min of fuel is given by:

0.05
Qcch  Q p kW
60 M

Figure 8.12 shows that the calculated rate of heat transfer from the combustion chamber
(Heat) is 6.98 kW. The corresponding value given by Cengel and Boles [1] is 6.89 kW.

8.6. Using macros in combustion analyses


Suppose that you want to investigate the effect of the amount of excess air (γ) on the
adiabatic flame temperature (Taf) of the three fuels: isooctane (C8H18), propane (C3H8),
and methane (CH4). You can do that by using the general Excel sheet for first-law
analyses and Goal Seek to determine the adiabatic temperature for C8H18 at different
values of γ, say, 0%, 10%, 20%, 30%, 40%, and 50%, and then plot Taf versus γ. Similarly,
the sheet can be used to determine the effect of excess air on the adiabatic flame
temperature for propane and methane. This way you have to repeat the procedure of
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 167

determining the adiabatic temperature by using Goal Seek 18 times for the three fuels
and six values of γ. Apart from being time-consuming, this procedure may also lead to
erroneous results due to human errors. Fortunately, Excel offers another useful feature
for dealing with such repetitive calculations and sensitivity analyses by creating macros.
The macro records the process of determining Taf at different amounts of excess air for
isooctane so that the procedure can be repeated for the two other fuels by activating the
macro. Figure 8.13 shows a macro named “Run” created to calculate the adiabatic flam
temperature for the three fuels at the six values of γ at the conditions of the automobile
engine considered in Example 8.5.

Figure 8.13. Excel sheet developed for the sensitivity analysis with a macro

After inserting the name of the fuel in cell C3, activating the macro will cause Excel to
determine the corresponding values of the adiabatic temperature and store them at a pre-
set location on the sheet. The results of the three fuels for Taf versus γ can be copied and
plotted as shown in Figure 8.14.

3000
C8H18
Adiabatic temperature (K)

2800 C3H8
CH4
2600

2400

2200

2000
0 10 20 30 40 50
Excess air (%)
Figure 8.14. Effect of excess-air on adiabatic flame temperature
168 Mohamed M. El-Awad

For all three fuels, Figure 8.14 shows that the adiabatic flame temperature decreases with
increasing percentage of excess air. The figure also shows that the adiabatic flame
temperature for methane is lower than that for propane, which is lower than that for
isooctane. This clearly indicates that the adiabatic flame temperature of a hydrocarbon
fuel increases with its molar mass.

8.7. Closure
This chapter illustrated the use of Thermax property functions for developing a general
Excel sheet for combustion analyses. The required input data for the sheet include the
fuel type, amount of excess air, air humidity, fuel inlet temperatures, air inlet temperature
and pressure, and the products exit temperature. A number of examples were given in
which the sheet was used to determine the dew-point temperature for the combustion
products and the amount of heat released from the combustion process. Comparisons with
solutions given in standard and advanced thermodynamics textbook proves the adequacy
of sheet for combustion analyses of both conventional and alternative fuels. An important
advantage of the sheet compared to the traditional use of property tables is that it allows
the use of Excel’s tools for “What-if” analyses like Goal Seek and Solver. The chapter
shows how the sheet can be used to deal with iterative combustion analyses such as
determining the adiabatic flame temperature of fuels.

Another useful feature offered by Excel for dealing with repetitive calculations and
sensitivity analyses of combustion processes is by creating macros. Appendix D extends
the sheet to perform additional first-law analyses of combustion processes, such as those
in closed-systems, while Appendix E extends it to deal with second-law analyses of
combustion processes. The general Excel sheet for combustion analyses is also useful
for developing Excel-based simulation models for internal-combustion engines [5]. By
taking into consideration the geometrical characteristics and ignition timing of the
engine, the models are more realistic than air-standard cycles and can be used to
investigate the effects of different design parameters on the engine’s performance.

References
[1] Y. A. Cengel and M.A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007
[2] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley & Sons, Inc, 2006.
[3] R.E. Sonntag, C. Borgnakke and G.J. van Wylen, Fundamentals of
Thermodynamics, 6th edition, John Wiley & Sons, Inc, 2003.
[4] W.W. Pulkrabek, Engineering Fundamentals of the Internal Combustion Engine,
Second Edition, Pearson Prentice Hall, 2004
[5] M.M. El-Awad, Developing simulation models with Excel for cooperative project-
based learning of I.C. engines.
https://www.academia.edu/82524318/Developing_Simulation_Models_with_Exc
el_for_Cooperative_Project_Based_Learning_of_I_C_Engines
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 169

Chapter 9

Analyses of vapour-absorption
refrigeration systems
170 Mohamed M. El-Awad

Compared to the more commonly used vapour-compression refrigeration systems, the


vapour-absorption refrigeration (VAR) systems require considerably less compression
work because they dissolve the refrigerant in a liquid solution at the lower pressure and
use heat to release it at the higher pressure. Therefore, VAR systems are important
elements for efficient use of energy as well as the utilisation of renewable energies. This
chapter focusses on the verification of Thermax functions for the analyses of VAR
systems. Thermax provides two sets of functions for the analyses of ammonia-water and
water-lithium-bromide VAR systems. The chapter verifies these functions by comparing
their calculations for single-effect VAR systems with relevant data obtained from the
published literature. The chapter then shows how the functions can be used to perform
energy analyses of a double-effect water-lithium-bromide VAR system for which
ASHRAE [1] gave some data for comparison.

9.1. Analyses the single-effect VAR system


Figure 9.1 shows a schematic diagram of a single-effect VAR system. This system
reduces the pumping work required to raise the pressure from that of the evaporator to
that of the condenser by dissolving the refrigerant in a liquid before compression in the
absorber. To free the refrigerants from the absorbent in the generator, the system needs a
low-temperature supply of heat. Here comes the other advantage of the system which is
that the required heat can come from the waste energy of the exhaust gas in power-
generation plants or from a renewable source of energy such as solar energy and biomass.
For more information about VAR systems, the reader can refer to ASHRAE [1] or
Stoecker and Jones [2].

Figure 9.1. A single effect VAR system (adopted from ASHRAE [1])
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 171

9.1.1. The analytical model


The coefficient of performance of a VAR system is defined as:

Qe
COP  (9.1)
Qg  w p

Where Qe is heat absorbed in the evaporator, Qg the heat provided to the generator
(desorber), and wp is the pump work. In terms of flow rates and fluid enthalpies, these
quantities are given by:

Qe  m 9 h10  h9  (9.2)

Qg  m 3h3  m 4h4  m 7h7 (9.3)

wp  m1h2  h1  (9.4)

The heat rejected in the absorber is given by:

Qa  m1h1  m6 h6  m10 h10 (9.5)

The solution concentrations X1 and X4 can be determined based on the given temperatures
Ta and Tg.

Calculation of flow rates:

m3  m2  m1 (9.6)
X 3  X 2  X1 (9.7)
X
m 4  m3 3
X4 (9.8)
m6  m5  m4 (9.9)
m7  m3  m4 (9.10)
m10  m9  m8  m7 (9.11)

Determination of temperatures:

T2 = T1 (9.12)

T4 = Tg (9.13)
172 Mohamed M. El-Awad

T5 = T4 – ε(T4 – T2) (9.14)

Where ε is the effectiveness of the solution heat-exchanger. The temperatures T3 and T6


can be determined based on the calculated solution concentrations.

Determination of enthalpies:
The enthalpies h1 , h4 , h5 at the given or calculated temperatures and concentrations can
be determined form the properties of the water-lithium bromide solution using the
relevant Thermax property functions and the enthalpies h7 , h8 , h10 at the given or
calculated temperatures can be determined form the properties of the water.

Enthalpy values at the remaining 4 points can be obtained as follows:

h2  h1 (9.15)

h3  h2 
m4
h4  h5  (9.16)
m2
h6  h5 (9.17)
h9  h8 (9.18)

In the following analyses the above model is applied for two single-effect VAR systems
that use ammonia-water and water-lithium bromide solutions.

9.1.2. Analysis of the ammonia-water VAR cycle


Sun [5] analysed the performance of the single-effect VAR system shown in Figure 9.1
with ammonia-water as the working fluid. For his case;

Tg=100oC, Tc=30oC, Ta=25oC, Te=-5oC,

The refrigerant mass flow rate was 1 kg/min, and the heat exchanger effectiveness was
80%.

The same system was simulated by using Thermax functions that determine the properties
of ammonia-water solution and Figure 9.2 shows the Excel sheet prepared for this
analysis. The given data are shown in the left side of the sheet while the main output
parameters are shown on the right side of it. The calculations part determines the
evaporator and condenser pressures as those given by the saturation pressures of
ammonia at the given temperatures. The calculations part then determines the values of
concentrations, mass flow rates, temperatures, and enthalpies at the different points in the
cycle. The results part determines the various rates of heat and work transfer in the cycle
before calculating the COP. The formulae used in the sheet are revealed by Figure 9.3.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 173

Figure 9.2. Excel sheet developed for the ammonia-water VAR system

(a)

(b)
Figure 9.3. The formulae in the Excel sheet of Figure 9.2 for the VAR system
174 Mohamed M. El-Awad

The figures on Table 9.1, which shows the cycle parameters as obtained by the present
add-in, show good agreements between the estimations of the present add-in and those
given by Sun [5].

Table 9.1. Verification of the add-in functions for ammonia-water VAR system
Sun [5] Thermax Error (%)
Qg (kW) 30.1314 30.86 2.42
Qc (kW) 18.4611 19.07 3.30
Qe (kW) 18.5974 18.58 -0.09
Qa (kW) 30.3269 30.38 0.18
w (kW) 0.0592 0.059 -0.34
Qex (kW) 11.8382 11.86 0.18
COP 0.6160 0.601 -2.44

9.1.3. Analysis of the water-lithium bromide VAR cycle


Abdulateef et al [8] analysed the performance of a water-lithium bromide VAR system
with the following data:

Te = 5oC, Tc = 35oC, Ta = Tc, Tg = 70-90oC

m1=0.016 kg/s

Heat exchanger effectiveness (ε) = 0.7

Mechanical efficiency of the pump (ηp) = 0.8

Figure 9.4 shows the Excel sheet developed for this Example. The left side of the sheet
shows the given data, which includes additional information about the pump efficiency.
The right side of the sheet shows the main output parameters. The calculations part
determines the evaporator and condenser pressures as those given by the saturation
pressures of water at the given temperatures. The calculations part determines the values
of concentrations, mass flow rates, temperatures, and enthalpies at the different points in
the cycle. The results part determines the various rates of heat and work transfer in the
cycle before calculating the COP. Figure 9.5 reveals the formulae included in the sheet
with Themax property functions for the water-lithium bromide solution.

The particular case shown in Figure 9.4 is the case where Tg = 80oC for which the COP
is 0.761. Table 9.2 and Table 9.3 compare the temperatures and enthalpy values at the
different point in the cycle, the rates of heat and work transfer, and the COP obtained by
the present model with their values given by Abdulateef et al [8].
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 175

Figure 9.4. The Excel sheet developed for the water-lithium bromide VAR cycle

Figure 9.5. The formulae in the Excel sheet of Figure 9.4 for the VAR system

Table 9.2. Comparison of temperatures and enthalpies


Point Temperature oC Enthalpy (kJ/kg)
Ref [8] Thermax Ref [8] Thermax
1 35 35 85.331 84.070
2 35 35 85.331 84.070
3 62 59.2 140.314 139.451
4 80 80 195.808 194.210
5 48.5 45.5 135.625 133.712
6 48.5 45.5 135.625 133.712
7 80 80 2650.317 2651.317
8 35 35 146.772 146.640
9 5 5 146.772 146.640
10 5 5 2510.452 2510.10
176 Mohamed M. El-Awad

Table 9.3. Comparison of heat and work transfer and COP at Tg = 80oC
Ref [8] Thermax Error (%)
Qe [kW] 3.147 3.2 1.68
Qc [kW] 3.328 3.39 1.86
Qa [kW] 3.962 4.01 1.21
Qg [kW] 4.151 4.2 1.18
Wp [kW] 0.000059 0.000059 0.00
COP 0.756 0.761 0.66

Figures 9.6 and 9.7 compare the estimations of the present add-in for the heat removed
in the absorber (Qa), the heat added in the absorber generator (Qg), the heat added in the
evaporator (Qe), the heat removed in the condenser (Qc), and the COP with those obtained
by Abdulateef et al [8].

8 8
7 Ref [8] 7 Ref [8]
6 Thermax 6 Thermax
5 5
Qg (kW)
Qa (kW)

4 4
3 3
2 2
1 1
0 0
60 70 80 90 100 60 70 80 90 100
Generator temperature (oC) Generator temperature (oC)
(a) (b)
6 7
Ref [8] 6 Ref [8]
5
Thermax 5 Thermax
4
Qe (kW)

Qc (kW)

4
3
3
2
2
1 1
0 0
60 70 80 90 100 60 70 80 90 100
Generator temperature (oC) Generator temperature (oC)
(c) (d)
Figure 9.6. Variation of (a) evaporator heat, (b) condenser heat, (c) absorber heat and
(d) generator heat with generator temperature
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 177

0.9
0.8
0.7
0.6
0.5

COP
0.4
0.3
Ref [8]
0.2
0.1 Thermax
0
60 70 80 90 100
Generator temperature (oC)
Figure 9.7. Variation of COP with generator temperature

Unlike the heat transfer in the evaporator (Qe) and condenser (Qc) where the fluid is pure
water, estimates of both Qa and Qg depend on the accuracy of estimating the enthalpy of
the water-lithium bromide solution. The figures show that the add-in functions give
accurate estimations of both Qg and Qa as well as the COP.

9.2. Analysis of a double-effect VAR system


Although VAR systems require less compression work than VCR systems, the COP of a
single effect VAR system is significantly lower than that of a VCR system. Therefore,
when the temperature of the available heat source is reasonably high, a double-effect
VAR system can be used. Figure 9.8 shows a schematic diagram of such a system in
which the heat rejected in the high-pressure condenser is utilized to provide the required
heating for the low-pressure desorber (generator).

Figure 9.8. Parallel-flow double effect water-lithium bromide VAR system (adapted
from ASHRAE [1])
178 Mohamed M. El-Awad

The case considered for verifying Thermax functions for the analyses of double-effect
VAR cycles has the following particulars given by ASHRAE [1]:

Capacity (Qe): 1760 kW


Solution: water-lithium bromide solution
Evaporator temperature (T10)= 5.0°C
Desorber solution exit temperature (T14):170.7°C
Condenser/absorber low temperature (T1 = T8): 42.4°C
Solution heat exchanger effectiveness (ε): 0.6

ASHRAE [1] provided data for this double-effect system case in which the temperature
in the low-pressure desorber (T4) is 97.8oC. This data is used for comparison with the
estimations of the present Thermax functions. Figure 9.9 shows the Excel developed for
the analysis sheet and Figure 9.10 reveals the formulae used in it.

Figure 9. 9. The sheet developed for analysing the double-effect VAR systems

Table 9.4 shows the values obtained by the Excel-Thermax model for the mass flow rate
(m ), refrigerant concentration (X), solution temperature (T), and solution enthalpy (h).
Figure 9.11 shows the percentage deviation of these results from those given by
ASHRAE [1]. The figure shows that most of the values deviated by less than 1% from
their corresponding values given by ASHRAE [1]. Table 9.5 compares the different
pressures and heat-transfer rates in the system’s components as obtained by the present
model with those given by ASHRAE [1]. The table, which also shows the pumps work
and cycle COP, shows excellent agreements with the reference data. The calculated COP,
in particular, is in close agreement with that given by the reference. Figure 9.12, that
shows the variation of the cycle's COP with the temperature of the low-pressure desorber,
reveals that the optimum COP is obtainable at temperatures of the low-pressure desorber
(T4) in the range 96-103oC.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 179

(a)

(b)
Figure 9.10. The formulae used for analysing the double-effect VAR systems
180 Mohamed M. El-Awad

Table 9.4. Thermax analysis of the double-effect VAR cycle


# m X T h
1 9.50 59.04 42.4 115.41
2 9.50 59.04 42.4 115.41
3 9.50 59.04 75.64 180.46
4 8.74 64.13 97.8 245.03
5 8.74 64.13 58.83 174.36
6 8.74 64.13 52.76 174.36
7 0.32 0.00 85.75 2661.97
8 0.75 0.00 42.4 177.57
9 0.75 0.00 5.0 177.57
10 0.75 0.00 5.0 2510.10
11 5.44 59.04 85.75 200.27
12 5.44 59.04 85.75 200.27
13 5.44 59.04 136.72 300.19
14 5.01 64.16 170.7 377.45
15 5.01 64.16 110.88 268.86
16 5.01 64.16 97.86 268.86
17 0.43 0.00 155.85 2788.36
18 0.43 0.00 102.8 430.99
19 0.43 0.00 42.38 430.99

0.5

0
10
1
2
3
4
5
6
7
8
9

11
12
13
14
15
16
17
18
19

-0.5
Deviation (%)

-1

-1.5

-2 m x

h T
-2.5

Figure 9.11. Deviations from ASHRAE [1] data


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 181

Table 9.5. Verification of Thermax estimations


ASHRAE[1] Thermax % error
Pa 0.88 0.870 -1.14
Pgl 8.36 8.441 0.97
Pgh 111.8 112.329 0.47
Qa 2328 2323.699 -0.18
Qcg 1023 1023.999 0.10
Qc 905 905.597 0.07
Qgh 1472 1469.296 -0.18
Qshx1 617 619.673 0.43
Qshx2 546 545.994 0.00
Wp1 0.043 0.043 -0.51
Wp2 0.346 0.341 -1.36
COP 1.195 1.198 0.21

1.3

1.2

1.1
COP

0.9

0.8
85 90 95 100 105 110 115

Temperature of low-pressure generator (oC)

Figure 9.12. Variation of COP with the low-pressure desorber temperature

9.3. Closure
The chapter presented first-law analyses of single-effect and double effect vapour-
absorption refrigeration cycles using Thermax property functions. For the single-effect
VAR system the results obtained with the two binary solutions, ammonia-water and
water-lithium-bromide, were verified by comparison with the data published by Sun [5]
and by Abdulateef et al [8]. For the double-effect systems the functions were compared
with the data provided by ASHRAE [1]. The analyses confirm the accuracy of Thermax
functions for all three cases considered.
182 Mohamed M. El-Awad

A current limitation of the add-in with respect to the analyses of vapour-absorption


refrigeration systems is the lack of functions that determine the chemical exergy of VAR
solutions. Research applications of the platform may also require additional vapour-
absorption solutions to be included in the library.

References
[1] American Society of Heating, Refrigeration and Air-Conditioning Engineers,
Handbook of fundamentals, Atlanta: ASHRAE 2005.
[2] W.F. Stoecker and J.W. Jones, Refrigeration And Air Conditioning, 2nd edition,
McGraw Hill Inc. 1982.
[3] J. Patek and J. Klomfar, "Simple functions for fast calculations of selected
thermodynamic properties of the ammonia-water system". International Journal of
Refrigeration, Vol. 18, No. 4, 1995, pp. 228 -234.
[4] M.R. Patterson and H. Perez-Blanco, “Numerical fits of the properties of lithium-
bromide water solutions”, ASHRAE Trans. Volume 94, Issue 2, pp. 2059-2077, 1988
[5] D-W Sun. "Comparison of the performances of NH3-H2O, NH3-LiNO3 and NH3-
NaSCN absorption refrigeration systems". Energy Convers. Mgmt, Vol. 39, No. 5/6,
1998, pp. 357-368.
[6] M.A.I. El-Shaarawi, S.A.M. Said, M.U. Siddiqui, "New correlation equations for
ammonia-water vapor liquid equilibrium (VLE) thermodynamic properties",
ASHRAE Transactions, 2013, pp. 293-298.
[7] P. Bourseau and R. Bugarel, “Absorption-diffusion machines: comparison of the
performances of NH3-H2O and NH3-NaSCN”. International Journal of
Refrigeration, Vol. 9, Issue 4, pp. 206-214.
[8] J.M. Abdulateef, M.A. Alghoul, R. Sirwan, A. Zahrim, K. Sopian, "Second law
thermodynamic analysis of a solar single-stage absorption refrigeration system",
Models and Methods in Applied Sciences, 2012 pp. 163 – 168.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 183

Appendices
184 Mohamed M. El-Awad

Appendix A: Ideal gases, refrigerants, and chemical reactants supported by


Thermax

A: Ideal gases
Table A.1 lists 29 gases included in the gas-group (Gas) with their respective reference
names used by Thermax. Values of the coefficients used to determine the different
properties of ideal gases are based on those given by Cengel and Boles [1].

Table A.1. Gases supported by the ideal-gas group (Gas)


# Gas Formula Thermax name
1 Nitrogen N2 N2
2 Oxygen O2 O2
3 Air — Air
4 Hydrogen H2 H2
5 Carbon monoxide CO CO
6 Carbon dioxide CO2 CO2
7 Water vapour H2O H2O
8 Nitric oxide NO NO
9 Nitrous oxide N2O N2O
10 Nitrogen dioxide NO2 NO2
11 Ammonia NH3 NH3
12 Sulphur S2 S2
13 Sulphur dioxide SO2 SO2
14 Sulphur trioxide SO3 SO3
15 Acetylene C2H2 C2H2
16 Benzene C6H6 C6H6
17 Methanol CH4O CH4O
18 Ethanol C2H6O C2H6O
19 Hydrogen chloride HCl HCl
20 Methane CH4 CH4
21 Ethane C2H6 C2H6
22 Propane C3H8 C3H8 or R290
23 n-Butane C4H10 C4H10n or R600
24 i-Butane C4H10 C4H10i or R600a
25 n-Pentane C5H12 C5H12
26 n-Hexane C6H14 C6H14
27 Ethylene C2H4 C2H4
28 Propylene C3H6 C3H6
29 Octane C8H18 C8H18

B: Vapour-compression refrigerants
Table A.2 lists the 28 refrigerants for VCR systems supported by the refrigerants group
(Ref) that include 18 halocarbon refrigerants, three inorganic refrigerants (ammonia,
water, and carbon dioxide), and seven hydrocarbon refrigerants included by ASHRAE
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 185

[2]. Currently, the group does not support the five cryogenic fluids. Chapters 5 and 6
illustrate the use of the functions in this group for the analyses of VCR systems.

Table A.2. Refrigerants supported by the refrigerants group (Ref)


# ASHRAE designation [2] Thermax M Tc [K] Pc [MPa] 
name [kg/kmol]
1 R-12 R12 120.91 385.12 4.136100 0.1795
(dichlorodifluoromethane)
2 R-22 (chlorodifluoromethane) R22 86.468 369.3 4.990000 0.2208
3 R-23 (trifluoromethane) . R23 70.0 299.29 4.832000 0.263
4 R-32 (methylenefluoride) R32 52.02 351.26 5.782000 0.2769
5 R-123 (dichlorotrifluroethane) R123 152.93 456.83 3.661800 0.2819
6 R-124 (2-chloro-1,1,1,2- R124 136.5 395.43 3.624300 0.2881
tetrafluoroethane)
7 R-125 (pentafluoroethane) R125 120.0 339.17 3.617700 0.3052
8 R-134a (tetrafluoroethane) R134a 102.03 374.21 4.059300 0.3268
9 R-143a (1,1,1-trifluoroethane) R143a 84.0 345.86 3.761000 0.2615
10 R-152a (difluoroethane) R152a 66.05 386.41 4.516800 0.2752
11 R-245fa (1,1,1,3,3- R245fa 134.05 427.16 3.651000 0.3776
pentafluoropropane)
12 R-1233zd (trans-1-chloro- R1233zd 130.5 438.75 3.580000 0.3414
3,3,3-trifluoroprop-1-ene)
13 R1234yf (2,3,3,3- R1234yf 114.0 367.85 3.382200 0.276
tetrafluoroprop-1-ene)
14 R1234ze (trans-1,3,3,3- R1234ze 114.04 382.52 3.636200 0.313
tetrafluoropropene)
15 R-404A [R-125/143a/134a R404A 97.604 345.2 3.729000 0.275
(44/52/4)]
16 R-407C [R-32/125/134a R407C 86.204 359.18 4.630000 0.3268
(23/25/52)]
17 R-410A [R-32/125 (50/50)] R410A 72.585 344.51 4.903000 0.296
18 R-507A [R-125/143a (50/50)] R507A 98.859 343.77 3.705000 0.2793
19 R-717 (ammonia) R717 17.02 405.4 11.333000 0.256
20 R-718 (water/steam) R718 18.0 647.1 22.064000 0.3443
21 R-744 (carbon dioxide) R744 44.0 304.1 7.380000 0.239
22 R-50 (methane) R50 16.044 190.56 4.599200 0.0114
23 R-170 (ethane). R170 30.07 305.32 3.872200 0.0995
24 R-290 (propane) R290 44.097 369.89 4.251200 0.1521
25 R-600 (butane) R600 58.124 425.13 3.796000 0.201
26 R-600a (isobutane) R600a 58.124 407.81 3.629000 0.184
27 R-1150 (ethylene) R1150 28.05 282.35 5.041800 0.0866
28 R-1270 (propylene) R1270 42.08 365.57 4.664600 0.146
186 Mohamed M. El-Awad

C: Combustion and chemical reactions


Thermax provides three general functions for the analyses of combustion process and
chemical reactions. The three functions, called Chm_f1, Chm_f2, and Chm_f3, require
as input the intended property as well as the substance name. Table A.3 lists the 29 fuels
and reactive substances which are the subjects of the first function Chm_f1, while Table
A.4 lists the 20 substances supported by the second function Chm_f2. Table A.5 shows
the four properties associated with Chm_f1 and Table A.6 shows the nine properties
associated with Chm_f2 for a general substance of the chemical representation CαHβOγNδ.

Table A.3. List of common reactants and reaction products in Chm_f1 function
# Substance Phase Thermax name
1 Carbon Gas C
2 Hydrogen Gas H2
3 Nitrogen Gas N2
4 Oxygen Gas O2
5 Carbon monoxide Gas CO
6 Carbon dioxide Gas CO2
7 Steam Gas H2O
8 Water Liquid H2OL
9 Hydrogen peroxide Gas H2O2
10 Ammonia Gas NH3
11 Methane Gas CH4
12 Acetylene Gas C2H2
13 Ethylene Gas C2H4
14 Ethane Gas C2H6
15 Propylene Gas C3H6
16 Propane Gas C3H8
17 n-Butane Gas C4H10
18 n-Octane Gas C8H18
19 n-Octane Liquid C8H18L
20 n-Dodecane Gas C12H26
21 Benzene Gas C6H6
22 Methyl alcohol Gas CH3OH
23 Methyl alcohol Liquid CH3OHL
24 Ethyl alcohol Gas C2H5OH
25 Ethyl alcohol Liquid C2H5OHL
26 Oxygen1 Gas O
27 Hydrogen1 Gas H
28 Nitrogen1 Gas N
29 Hydroxyl Gas OH
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 187

Table A.4. List of fuels and reacting substances in Chm_f2 function


# Substance Formulae Thermax name
1 Gasoline C8H15 C8H15
2 Light diesel C12.3H22.2 LDL
3 Heavy diesel C14.6H24.8 HDL
4 Isooctane C8H18 C8H18
5 Methanol CH3OH CH3OH
6 Ethanol C2H5OH C2H5OH
7 Methane CH4 CH4
8 Propane C3H8 C3H8
9 Nitromethane CH3NO2 CH3NO2
10 Heptane C7H16 C7H16
11 Cetane C16H34 C16H34
12 Heptamethylnonane C12H34 C12H34
13 α-methylnaphthalene C11H10 C11H10
14 Carbon monoxide CO CO
15 Coal (carbon) C C
16 Butene-1 C4H8 C4H8
17 Triptane C7H16 C7H16
18 Isodecane C10H22 C10H22
19 Toluene C7H8 C7H8
20 Hydrogen H2 H2

Table A.5. Properties provided by the first function, Chm_f1


Property [unit] Thermax name
1 Molar mass (M), [kg/kmol] M
2 0
Molar enthalpy of formation ( h f ), [kJ/kmol] hf0
3 Molar Gibbs function of formation ( g 0f ), [kJ/kmol] gf0
4 Molar absolute entropy ( s 0 ), [kJ/kmol.K] as0

Table A.6. Properties provided by the second function, Chm_f2


Property [unit] Thermax name
1 Molar mass (M), [kg/kmol] M
2 Gravimetric higher heating values (QH), [kJ/kg] QH
3 Gravimetric lower heating value (QL), [kJ/kg] QL
4 Heat of vaporisation (Hv), [kJ/kg] HOV
5 Stoichiometric air-fuel ratio (AFs) AF
6 Carbon content of the fuel (α) ALFA
7 Hydrogen content of the fuel (β) BETA
8 Oxygen content of the fuel (γ) GAMA
9 Nitrogen content of the fuel (δ) DELTA
188 Mohamed M. El-Awad

Note that for Chm_f1, steam is referred to by “H2O”, while liquid water is referred to by
“H2OL”. The respective properties for the fuels and reactive substances listed in Table
A.3 and Table A.4 are based on those given by Cengel and Boles [1] and Pulkrapek [3],
respectively. Case 7 in Table 2.1 shows how the function Chm_f1 can be used to
determine the molar mass (M) of benzene (C6H6).

The third function, Chm_f3, returns the values of the specific heat, enthalpy, internal
energy, and temperature-dependent entropy change of the ideal gases listed in Table A.1
on a unit-mole basis. Table A.7 shows the four properties for ideal gases associated with
Chm_f3 expressed on a unit-mole basis. Chapter 8 illustrates the use of the three
functions in this group for the analyses of combustion processes.

Table A.7. Properties provided by the third function, Chm_f3


Property [unit] Thermax name
1 Molar specific heat ( c p ), [kJ/kmol.K] cpm
2 Sensible molar internal energy ( u ), [kJ/kmol] um
3 Sensible molar enthalpy ( h ), [kJ/kmol] hm
4 Temperature-dependent molar entropy ( s 0 ), [kJ/kmol.K] s0m

References
[1] Y. A. Cengel and M.A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[2] American Society of Heating, Refrigeration and Air-Conditioning Engineers,
(ASHRAE), Handbook of fundamentals, Atlanta, 2017.
[3] W.W. Pulkrabek, Engineering Fundamentals of the Internal Combustion Engine,
Second Edition, Pearson Prentice Hall, 2004.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 189

Appendix B: Thermodynamic properties of ideal gases and humid air

The property functions provided by Thermax for the ideal-gases group (Gas) and the
humid-air group (Psy) are based on established thermodynamic relationships [1].
Psychrometric analyses also require certain properties of liquid water. This appendix
reviews the relationships used to develop the functions in these two add-in groups.

A. Properties of ideal gases


Properties of ideal gases at a given temperature are obtained by suitable integration of
~
the following equation for the molar specific heat at constant pressure ( c p ):

c~p  a0  a1T  a 2T 2  a3T 3 [kJ/kmol.K] (B.1)

Where T is the absolute temperature and a0, a1, a2, and a3 are constants that have different
values for the 29 gases listed in Table A.1.

~
Accordingly, the molar enthalpy ( h ), molar internal energy ( u~ ), and temperature-
~0
dependent molar entropy change ( s ) of an ideal gas are calculated from [1]:

T
~ ~
h  h0   c~p (T )dT (B.2)
T0
T 2
u~  u~0   c~v (T )dT  u~0   (c~p (T )  Ru )dT (B.3)
T0 1
T ~
~s 0  ~s 0  c p (T ) dT
0  T
T0
(B.4)

~
Where, h0 , u~0 , and ~
s 00 are specified values at a reference temperature (T0) and Ru is the
~
universal gas constant. For h and u~ , the reference temperature is taken as 300K and the
~
corresponding values of h0 and u~0 are those given by Cengel and Boles [1]. However,
for ~s 0 the reference temperature is taken as 298K and the corresponding value of ~s 0 is
0
the absolute entropy of the gas. Thermax functions do not return the molar properties of
the ideal-gases as given above, but per kg, i.e., h, u and s0, given by:

~
hh /M (B.5)

u  u~ / M (B.6)
190 Mohamed M. El-Awad

s 0  ~s 0 / M (B.7)

Where M is the molar mass of the gas. The relative pressure (Pr) and relative volume (vr)
of the ideal gases, respectively, are obtained from [1]:

Pr = exp(s°/R) (B.8)

v r  T / Pr (B.9)

Table A.1 shows the respective add-in reference names for the 29 gases. Chapter 3
illustrates the use of the Gas functions for the analyses of gas-power cycles.

B. Properties of humid air


For psychrometric analyses the enthalpy of humid air (h) is calculated from [1]:

h  c paT  hv (B.10)

Where, cpa is the specific heat at constant pressure for dry air (cpa =1.005 kJ/kg.K), T is
the temperature in oC, ω is the absolute humidity, and hv is the enthalpy of water vapour
at the air temperature and partial-pressure of the water-vapour (Pv). The absolute
humidity (ω) in Equation (B.10) is determined from:

  0.622 Pv / P  Pv  (B.11)

Where, P is the atmospheric pressure, which is total pressure of the air-water-vapour


mixture, and Pv is the partial pressure of water-vapour in the air. The partial pressure of
water vapour itself is determined from:

Pv  Psat / 100 (B.12)

Where ϕ is the relative humidity and Psat is the saturation pressure of water at the given
dry-bulb temperature. The enthalpy of water vapour (hv) in Equation (B.10) is
approximated by the following equation [1]:

hv  2500 .09  1.82T (B.13)

Where 2500.09 (kJ/kg) is the enthalpy of saturated water vapour at 0°C and 1.82 is the
average cp value of water vapour in the temperature range -10 to 50°C. Property functions
of this group determine the relative humidity from the following relationship:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 191

  P / 0.622   Pg (B.14)

This group needs functions that determine the saturation pressure (Psat) and saturation
temperature (Tsat) for water in the temperatures range 0 – 100oC usually met in common
air-conditioning practice.

For the convenience of making this group independent from the water-group, it has its
own functions (PsyPsat_T and PsyTsat_P) that determine these two properties from [2]:

B
A
Psat  0.1333  10 C T
(B.15)

  P 
Tsat  B /  A  log    C (B.16)
  0.1333  

Where, T is in oC, P in kPa and the three constants A, B, and C, respectively, have values
of 8.07131, 1730.63, 233.426 for 1< T < 100oC and 8.14019, 1810.94, 244.485 for 99
< T < 374oC.

A third auxiliary function in this group (PsyhL_T) determines the enthalpy of saturated
liquid water at a given temperature (hL). This is obtained from the following equation
which was obtained by curve-fitting the data for water in the range 0 – 100oC using
Excel’s trendline feature:

hL  0.146  4.184 T (B.17)

This function that uses equation (B.17) also makes the Psy group independent from the
Wat group. Chapter 7 illustrates the use of the Psy functions for the analyses of typical
air-conditioning processes and systems.

References
[1] Y. A. Cengel and M. A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[2] Wikipedia contributors, “Antoine equation”, Wikipedia, The Free Encyclopaedia,
internet, https://en.wikipedia.org/wiki/Antoine_equation (Last accessed
November 23, 2015).
192 Mohamed M. El-Awad

Appendix C: Verification of Thermax functions for superheated refrigerants

Thermax functions for saturated refrigerants interpolate the data provided by ASHRAE
[1], but its functions for superheated refrigerants apply approximate formulae. While the
function for the specific volume of superheated refrigerants uses the Soave-Redlich-
Kwong (SRK) equation, the functions for enthalpy and entropy use approximate constant
specific-heat relationships. This appendix verifies the accuracy of these functions by
comparing their estimations with those given by ASHRAE [1] for refrigerant R134a.

A. Formulae for the properties of superheated refrigerants


Specific volume
The molar specific volume ( v~ ) of a superheated refrigerant is obtained from the Soave-
Redlich-Kwong (SRK) equation of state [2]:

RT a
P= ~u ~ ~ (C.1)
v  b v (v  b)

where, Ru is the universal gas constant, P is the absolute pressure, and T is the absolute
temperature. The constants a, b and , which depend on the refrigerant’s pressure and
temperature at the critical point, are given by:

a = 0.4278 Ru2 Tc 2 / Pc (C.2)

b = 0.0867 Ru Tc / Pc (C.3)

 
  1  S 1  Tr 
2
(C.4)

Where, Tc and Pc are the temperature and pressure at the critical point, Tr =T/Tc is the
reduced temperature, and S is the following function for the acentric factor ():

S = 0.48508 + 1.55171   0.15613 2 (C.5)

Values of the acentric factor itself are mostly obtained from ASHRAE [1] or other
relevant sources. It can also be calculated from [3]:

 
   log 10 prsat  1 , at Tr = 0.7 (C.6)

Where Prsat =Psat/Pc is the reduced saturation pressure. Values of Tc, Pc, and  for the 28
refrigerants supported by Thermax are shown in Table A.2.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 193

Equation (C.1), which is a non-linear equation that requires a numerical solution, is


solved by using a Newton-Raphson solver provided by Thermax. The specific volume
(v), in m3/kg, is then obtained from the molar specific volume, v~ , as follows:

v  v~ / M (C.7)

Where M is the molar mass of the refrigerant and its values for the 28 refrigerants
supported by Thermax are shown in Table A.2.

Enthalpy
Enthalpy (h) of a superheated refrigerant at a given pressure and temperature is
determined from the following relationship:

h  h g  Cp *g T  T s  (C.8)

Where hg and Ts are the enthalpy and temperature of the saturated vapour refrigerant at
the given pressure, P, while Cp *g is the value of its specific heat evaluated at a skewed
pressure (P*) given by:

P*  f  P (C.9)

Where f is an adjustment factor that can be tuned to improve the accuracy of Equation
(C.8). The following validation analysis shows that this can be achieved by making f a
function of the ratio of the actual pressure to the fluid’s critical pressure. The analysis
also shows that f = 0.5 usually gives a good accuracy. Values of hg, Ts, and Cp *g in
Equation (C.8) are obtained from ASHRAE data for saturated refrigerants [1].

Entropy
Entropy (s) of a superheated refrigerant at a given pressure and temperature is determined
from the following relationship:

Cp *g T  Ts 
s  sg  (C.10)
Tav  273

*
Where sg is the entropy of saturated vapour refrigerant at the given pressure, Cp g ,is the
value of the specific heat evaluated at the skewed pressure (P*) given by Equation (C.9),
and Tav is an average temperature calculated as follows:

Tav  T  Ts  / 2 (C.11)
194 Mohamed M. El-Awad

The temperature at a given pressure and enthalpy or entropy


There are two functions in this subgroup that determine the temperature (T) of a
superheated refrigerant given its pressure (P) and its enthalpy (h) or entropy (s). To
determine T from P and h, Equation (C.8) is rearranged as follows:

 
T  Ts  h  h g / Cp *g (C.12)

Where Ts and hg are values of the saturation temperature and enthalpy of saturated
refrigerant vapour at the given pressure, but Cp *g is the value of the specific heat of
saturated refrigerant vapour determined at the skewed pressure P*. Similarly, when the
pressure and entropy of the superheated refrigerant are known and its temperature is to
be determined, the following equation is obtained from Equation (C.10):

ssg


T  Tg  273 e  Cp*g
 273 (C.13)

Where Tg and sg are the temperature and entropy of saturated vapour refrigerant at the
*
given pressure, while Cp g is the value of the specific heat of saturated refrigerant vapour
determined at the skewed pressure P*.

B. Verification of the functions for the properties of superheated refrigerants


Figure C.1 compares the densities for refrigerant R134a calculated by using Thermax
function, Refv_PT, with those given by ASHRAE [1] for R134a at pressures of 0.1 MPa,
1.4 MPa, and 3.0 MPa. The figure confirms the accuracy of the estimations obtained by
the function at all three pressure levels. Figure C.2 compares the values of enthalpy
determined by Thermax function Refh_PT on the basis of Equations (C.8) and (C.9) with
the corresponding values given by ASHRAE [1] at the same pressure levels. The Figure
shows that the values obtained with f= 1.0 deviate considerably from ASHRAE data,
especially at P = 3.0 MPa. Equation (C.8) leads to a good accuracy at P 3.0 MPa by using
f= 0.5, but not at P = 0.1 or 1.4 MPa.

As mentioned earlier, the accuracy of Equation (C.9) can be improved by making the
factor f a function of the pressure ratio to the critical pressure. In this respect, two options
were investigated the first of which used the following formula that was obtained by
fitting the computed enthalpy values for R134a to ASHRAE data at the three pressures:

 1.55 P 
 
 Pc 
f  1.3  e 
(C.14)

Where Pc is the critical pressure (for R134a Pc = 4.0593 MPa).


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 195

5.5 72.5
ASHRAE ASHRAE
5 62.5
Density (kg/m3)

Thermax Thermax

Density (kg/m3)
4.5
52.5
4
42.5
3.5
3 32.5
P=0.1 MPa P=1.4 MPa
2.5 22.5
-50 0 50 100 150 200 0 100 200 300
Temperature (oC) Temperature (oC)
(a) (b)
150
ASHRAE
130
Density (kg/m3)

Thermax
110

90

70
P=3.0 MPa
50
50 150 250 350
Temperature (oC)
(c)
Figure C.1: Accuracy of the computed densities for R134a at: (a) 0.1 MPa, (b) 1.4 MPa,
and (c) 3.0 MPa

Note that according to Equation (C.14), the value of f converges to 0.276 as the pressure
approaches Pc and to a value of 1.3 at low pressures. The second pressure-dependent
function tried for f makes it equal to the compressibility factor (Z), i.e.:

f  Pv g / RTs (C.15)

Where Ts and vg are the saturation temperature and specific volume of the saturated
vapour at the given pressure, respectively, and R is the gas constant. The value of vg was
determined from ASHRAE data by using the relevant Thermax function. The enthalpy
values computed with f given by Equation (C.14) “f=exp” and by Equation (C.15) “f=Z”
are also shown in Figure C.2. The figure shows that both options give similar or better
accuracy than f = 0.5 at the three pressure levels, but Equation (C.14) is more accurate
than Equation (C.15) at P = 3.0 MPa. The largest deviations of the computed enthalpy
196 Mohamed M. El-Awad

values from ASHRAE data occur at P=0.1 MPa, which increase as the temperature
increases, i.e. when the gas tends to behave more ideally.

600 700
P=0.1 MPa P=1.4 MPa
650
550

Enthalpy (kJ/kg)
600
Enthalpy (kJ/kg)

500 550
500
450 ASHRAE ASHRAE
450
Thermax, z=1.0 Thermax, z=1.0
Thermax, z=0.5 400 Thermax, z=0.5
400 Thermax, z=exp Thermax, z=exp
350
Thermax, z=Z Thermax, z=Z
350 300
-50 0 50 100 150 200 50 150 250 350
Temperature (oC) Temperature (oC)
(a) (b)
1100
ASHRAE
1000 Thermax, z=1.0
Thermax, z=0.5
900
Enthalpy (kJ/kg)

Thermax, z=exp
800 Thermax, z=Z

700
600
500
400
P=3.0 MPa
300
50 150 250 350
Temperature (oC)
(c)
Figure C.2. The accuracy tests for different estimations of computed enthalpy values for
R134a at: (a) 0.1 MPa, (b) 1.4 MPa, and (c) 3.0 MPa

Figure C.3 compares the values of entropy determined by the function Refs_PT on the
basis of Equation (C.10) with the corresponding values given by ASHRAE for R134a at
the three pressure levels of 0.1, 1.4, and 3.0 MPa. The figure shows the computed entropy
values obtained with f=1.0, f=0.5, f given by Equation (C.14), and f given by Equation
(C.15). As the figure shows, the last three options are reasonably accurate, but best
agreement with ASHRAE data is obtained with f given by Equation (C.15). El-Awad et
al [4] verified Thermax functions for two halocarbon refrigerants (R410A and R1234yf)
and two natural refrigerants (R290, and R744). Chapters 5 and 6 illustrate the use of the
functions in this group for the analyses of simple, multi-stage compression, and cascade
VCR cycles.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 197

2.4 2.4
P=0.1 MPa P=1.4 MPa
2.2 2.2

Entropy (kJ/kg.K)
Entropy (kJ/kg.K)

2 2

1.8 1.8
ASHRAE ASHRAE
1.6 Thermax, z=1.0 1.6 Thermax, z=1.0
Thermax, z=0.5 Thermax, z=0.5
1.4 Thermax, z=exp 1.4 Thermax, z=exp
Thermax, z=Z Thermax, z=Z
1.2 1.2
-50 0 50 100 150 200 0 100 200 300
Temperature (oC) Temperature (oC)
(a) (b)
3.2
ASHRAE
Thermax, z=1.0
2.8 Thermax, z=0.5
Entropy (kJ/kg.K)

Thermax, z=exp
2.4 Thermax, z=Z

1.6
P=3.0 MPa
1.2
50 150 250 350
Temperature (oC)
(c)
Figure C.3. The accuracy tests for different estimations of computed entropy values
for R134a at: (a) 0.1 MPa, (b) 1.4 MPa, and (c) 3.0 MPa

References
[1] American Society of Heating, Refrigeration and Air-Conditioning Engineers,
(ASHRAE), Handbook of fundamentals, Atlanta, 2017.
[2] Y. A. Cengel and M. A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[3] Wikipedia contributors, “Acentric factor”, Wikipedia, The Free Encyclopaedia,
internet: https://en.wikipedia.org/wiki/Acentric_factor, (last accessed November 30,
2015.
[4] M.M. El-Awad, M.S. Al Nabhani, K.S. Al Hinai, A. Younis, Development and
Validation of an Excel Add-In for determining the Properties of Various
Refrigerants, Proceedings of the 1st National Conference on Applied Science,
Engineering & Technology 2019, CASET – 2K19, 11th June 2019, Ibri, Oman.
198 Mohamed M. El-Awad

Appendix D: Additional first-law analyses of combustion processes

This appendix shows how the general Excel sheet for combustion analyses developed in
Chapter 8 can be used to deal with additional standard combustion analyses. The analyses
considered in this appendix include determining a fuel’s enthalpy of combustion and
heating value and first-law analyses of combustion in closed systems.

D.1. Enthalpy of combustion and heating value of a fuel


Since the chemical compositions of fuels like coal, oil, or biogas vary widely, their
enthalpy of formation cannot be obtained easily from tables. Therefore, the analysis
method presented in the chapter 8 by using the enthalpy of formation is not suitable for
dealing with these fuels. A more appropriate method in this case is to use the fuel’s
enthalpy of combustion instead of its enthalpy of formation. The enthalpy of combustion
of a fuel is defined as the amount of heat released during a steady-flow combustion
process when 1 kmol (or 1 kg) of fuel is burned completely at a specified temperature
and pressure [1]. The enthalpy of combustion can be measured experimentally for any
fuel. Equation (8.33) can be rearranged in the following form:

N N N N
~0 ~0 ~ ~
Q W  P h fP  R h fR  P h P  R hR (D.1)
P R P R

The underlined terms in the right side of the above equation is nothing but the enthalpy
~
of combustion at the standard temperature and pressure ( hC0 ). Therefore, the above
equation can be written as:

N N
~ ~ ~
Q  W  hC0  P hP  R hR (D.2)
P R

If the combustion process involves no work transfer, Equation (D.2) reduces to:

N N
~ ~ ~
Q  hC0  P hP  R hR (D.3)
P R

Equation (D.3) indicates that the rate of heat-transfer from the combustion process can
~
be obtained by adding the fuels’ enthalpy of combustion at the standard state ( hC0 ) to the
difference in enthalpy deviations of the products and reactant from their respective values
at the standard state of 25oC, 1 atm. Thus, Equation (D.3) provides an alternative method
to determine the heat-transfer from a combustion process.

~
The positive value of hC0 is known as the heating value (HV) of the fuel. Depending on
the state of water in the products of combustions, two heating values for a given fuel are
quoted; which are the higher heating value (HHV) and the lower heating value (LHV).
The higher heating value is obtained when all the water formed by combustion is a liquid,
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 199

while the lower heating value is obtained when all the water formed by combustion is a
vapour. The two heating values differ by the amount of energy that would be required to
vaporise the liquid water in the products. Thermax function “Chm_f2” returns the higher
and lower heating values per kg for the fuels listed in Table A.4. The following example
illustrates how the heat of combustion can be determined by using Equation (D.3).

Example D.1. Combustion analysis using the enthalpy of combustion


Determine the amount of heat transferred in the spark-ignition engine considered in
Example 8.2 which operates on stoichiometric propane fuel (C3H8) using the heating-
value of propane.

Solution
Figure D.1 shows the Excel sheet developed for this example that uses the heating value
of propane instead of its enthalpy of formation. The formula window reveals the formula
in cell R4 that calculates the heat transfer (Q2_) according to Equation (D.3). The
calculated value for the heat transfer is -1,577,339 kJ/kmol of fuel. The value found in
Example 8.2 by using the data for enthalpy of formation is 1,577,317 kJ/kmol of fuel.

Figure D.1. Excel sheet developed for combustion analysis using the enthalpy of
combustion
~
The enthalpy of combustion at standard temperature and pressure ( hC0 ) can be determined
by using the data for enthalpy of formation:

N N
~ ~0 ~0
hC0  P h fP  R h fR (D.4)
P R

The following example illustrates the procedure of determining the heating value from
the enthalpy of formation data.
200 Mohamed M. El-Awad

Example D.2. Calculating enthalpy of combustion for methane


Calculate the enthalpy of combustion of gaseous methane, in kJ per kg of fuel, (a) at
25oC, 1 atm with liquid water in the products, (b) at 25oC, 1 atm with water vapour in the
products. (c) Repeat part (b) at 1000 K, 1 atm.

This example is adopted from Moran and Shapiro [1], Example 13.7.

Solution
(a) Figure D.2 shows the modified Excel sheet for part (a) of this example in which the
amount of excess air (γ) has been assigned a zero value and all the temperatures have
been assigned a value of 298K. The formula window reveals the formula in cell O11 that
determines the fraction of water vapour in the products (Nwv_ex).

Figure D.2. Excel sheet for determining the higher heating value for methane

As the formula window shows, this fraction has been forced to be zero so that all the
water in the products (2 moles) will be in the liquid phase. Dividing the heat of
combustion (Q_p = 890,330 kJ/kmol of fuel) by the molar mass for methane, which is
16.04, gives the heating value (HV). The calculated value, which is 55506.9 kJ/kg, agrees
with the higher heating value of methane shown in Table A.4 in Appendix A.

(b) Figure D.3 shows the sheet developed for this part in which all the water in the
products has been forced to be in the vapour phase by making the necessary change to
the formula in cell O11. The value of the heating value (HV) has now changed to 50,019.3
KJ/kg, which is the lower heating value for methane as given in Table A.4 in Appendix
A.

(c) Figure D.4 shows the general Excel sheet as modified for this part. Note that the
temperatures of the fuel, combustion air, and products of combustion are all assigned a
value of 1000K. When the combustion products are at a temperature of 1000K and
atmospheric pressure, all the water in the products will be in the vapour phase. The
heating value determined in this case is 49,920.0 kJ/kg. The value obtained by Moran
and Shapiro [1] is 49,910.0 kJ/kg.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 201

Figure D.3. Excel sheet for determining the lower heating value for methane

Figure D.4. Excel sheet for determining the heat value for methane at 1000K

For gaseous fuels (treated as ideal gasses), the heating values are usually stated on a
volumetric basis rather on a mass basis. If the heating value on a mole basis (HV) is
known, one can obtain that per cubic meter (HV´) from the following relation [2]:

HV
HV   [kJ/m3] (D.5)
24.5

The following example demonstrates the use of the general combustion sheet for
determining the volumetric heating values for fuel mixtures. The example is based on
Example 4 in Chapter 11 by Annamalai and Puri [2].

Example D.3. Determining the heating values for biogas


If biogas is assumed to consist of 60% methane and 40% carbon dioxide (volume %),
what is its higher heating value at 298K?

Solution
The sheet developed for this example requires some extensions to the general sheet
developed for determining the mass-based heat value of a single fuel. Figure D.5 shows
202 Mohamed M. El-Awad

these extensions that determine the heating value for CH4 per m3 when water vapour is
in the vapour phase. Note that formula window in Figure D.5 reveals the formula in cell
R6 that determines the volumetric heat value for CH4 according to Equation (D.5).

Figure D.5. Determining the higher volumetric heat value for biogas

The volumetric heating vale for biogas is given by:

'
HVbiogas  0.6HVCH
'
4  0.4HVC 2O
'

Since the heat value for CO2 is zero, the heat value for biogas is 60% of that of CH4. The
heat value thus determined, which is 21804 kJ per m3, is the same as that obtained by
Annamalai and Puri [2].

D.2. First-law analyses of closed system combustion


For combustion in a closed system, the first-law of thermodynamics reads:

Q W  N
P
PuP
~ 
N
R
~
RuR (D.6)

Where, u~ is the molar internal energy of the reactant or product of combustion. For ideal
gases:
~
u~  h  RuT (D.7)

Where, Ru = 8.314 kJ/kmol.K is the universal gas constant. Substituting in Equation


(D.6):

 N h   N h~ 
~
Q W  P P  Ru T P  R R  Ru T R (D.8)
P R

~ ~ ~
Replacing h by h f  h and rearranging:
0
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 203

 N h   N h~  N N
~ ~ ~
Q W  P
0
fP  h P  R
0
fR  h R  Ru T P P  Ru T R R (D.9)
P R P R

The underlined term on the right side of the equation is equal to the heat transfer in a
steady, constant-pressure combustion. Representing this heat transfer by (Qp), Equation
(D.9) can be written as:

 
Q  W  Q p  Ru  T P


N
P
P  TR N
R
R



(D.10)

The following examples illustrate the use of the general combustion sheet to determine
the heat-transfer (Q) and adiabatic flame temperature (Taf) when combustion takes place
in a closed system.

Example D.4. Closed-system heat transfer from the combustion of pentane


A closed, rigid vessel initially contains a gaseous mixture of 1 kmol of pentane (C 5H12)
and 150% of theoretical air at 25oC, 1 atm. If the mixture burns completely, determine
the heat transfer from the vessel, in kJ, and the final pressure in atm, for a final
temperature of 800 K.

This example is based on problem 13.40 given by Moran and Shapiro [1].

Solution
The general sheet developed for steady-flow combustion analyses can also be used for
closed-system combustion by adding a single cell in the results part to determine Qv from
Q according to Equation (D.10). The formula window in Figure D.6 reveals the formula
in this cell, which is cell R6. The calculated value for Qv, which is 2,561,006 kJ/kmol,
closely agree with the value given by Moran and Shapiro [1], which is 2,563,732 kJ/kmol.

Figure D.6. Excel sheet developed for closed-system combustion of pentane


204 Mohamed M. El-Awad

Example D.5. Adiabatic flame temperature for liquid octane in a closed system
Liquid octane enters a rigid closed system reactor with 40% excess air. Determine the
adiabatic flame temperature with both the fuel and air being at 298 K and 1 atm.

This example is based on Example 9 given by Annamalai and Puri [2] in Chapter 11.

Solution
The sheet for this example was developed by making the necessary modifications to that
of the previous example. Figure E.7 shows the modified sheet in which the fuel is
C8H18L and the amount of excess air is 40%.

Figure D.7. Adiabatic flame temperature of liquid octane in a constant volume device

Since the fuel is liquid, the fuel’s enthalpy change (ΔhR_fuel) in cell L2 is zero. Goal
Seek was used to find the adiabatic flame temperature, which is the exhaust temperature
(T_ex) at which Qv is zero. Figure D.7 shows the solution obtained, which is 2359.4K.
The solution obtained by Annamalai and Puri [2] is 2340K.

References
[1] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics,
5th edition, John Wiley & Sons, Inc, 2006.
[2] K. Annamalai and I.K. Puri. Advanced Thermodynamics Engineering, CRC 2002
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 205

Appendix E: Second-law analyses of combustion processes

This appendix extends the general Excel sheet developed in Chapter 8 for combustion
analyses to deal with second-law analyses that determine the rate of entropy production
in the combustion process and the maximum work that can be done during the process.

E.1. Second-law equations


The purpose of fuel combustion is to release the fuel’s chemical energy in the form of
heat as a necessary step towards converting it into useful work via a heat engine. In this
respect, first-law analyses of combustion reactions are useful for quantifying the amount
of thermal energy that can be obtained from the combustion process, but they give no
indication of the inherent irreversibility of combustion process that leads to significant
losses in the work potential of the fuel’s chemical energy. By applying the second law of
thermodynamics, we can determine the amount of available (reversible) work from the
fuel’s energy and rate of exergy destruction in the combustion process.

Exergy and reversible work in combustion processes


The property that measures the work potential of the energy content of a fluid stream is
its exergy χ (also called availability). Disregards mixing and chemical reactions, the
thermo-mechanical exergy is defined as:

~ ~

  h  h0  T0 ~s  ~s0  
V2
2
 gz (E.1)

Where, h0 and s0 are enthalpy and entropy of the stream at the dead state at (T0, P0),
respectively, V is the velocity, and z the elevation. Exergy values of the reactants and
products enable us to determine the reversible work Wrev, which is the maximum work
that can be done during a process. In the absence of any changes in kinetic and potential
energies, the reversible work relation for a steady-flow combustion process that involves
heat transfer with only the surroundings at T0 can be expressed as [1]:

 N h    N h~ 
~ ~ ~ ~ ~
Wrev  R
0
f  h  h 0  T0~
s R P
0
f  h  h 0  T0~
s P
(E.2)
R P

 N h    N h~ 
~ ~ ~
W rev  R
0
f  h  T0 ~
s R P
0
f  h  T0 ~
s P
(E.3)
R P

~ ~ ~ ~ ~
Note that the enthalpy term h  h 0 in Equation (E.1) has been replaced by h f0  h  h 0 ,
~
where h f0 is the enthalpy of formation at the standard reference state.

Exergy destruction and entropy production in the combustion process


The exergy destroyed ψdestroyed associated with a chemical reaction can be determined
from:
206 Mohamed M. El-Awad

 destroyed  T0 (E.4)

Where σ is the total entropy change or the entropy generation. In combustion processes,
entropy is produced by the chemical reaction itself and also as a result of heat-transfer to
the surroundings which are usually at a lower temperature. The entropy produced (σ) per
unit mole of fuel in a closed or steady-flow reacting system can be obtained from [1]:

T
Qi
  S p  SR  0 (E.5)
i i

Where SP and SR are the total entropy of the combustion products and reactants,
respectively, per mole of fuel and Qi is the portion of total heat transfer crossing the
section of the of temperature is Ti. Equation (E.5) adopts the usual sign convention that
the positive direction of heat transfer is into the system. In terms of entropies of the
reactants and products components, Equation (E.5) reads:

N N T
~  ~  Qi
 P sP R sR 0 (E.6)
P R i i

Where ~sP and ~sR are the absolute molal entropy of the components in the products and
reactants, respectively. For component i of a mixture of ideal gases, the absolute entropy
at any temperature T and partial pressure Pi can be obtained from [1]:

s T , Pi   ~
sio T , P0   Ru ln i m
~ yP
(E.7)
P0

Where yi is the molal fraction of the gas, Pm is the total pressure of the mixture, and P0 =
1 atm.

E.2. Extending the workbook for second-law combustion analyses


Thermax provides the function “Chms0m_1TK” that returns the absolute molal entropy
at P0 = 1 atm and any temperature T, which is the underlined term in Equation (E.7), for
all the gases supported by Thermax. Another function, “Chmas0_1”, returns the absolute
molal entropy at the standard temperature and pressure (T0 = 25oC, P0 = 1 atm) of the all
the substances supported by Thermax as listed in Table A.3. The following examples
extend the general Excel sheet for performing second-law analyses to determine the
reversible work of a combustion process and the rate of entropy production.

Example E.1. Second-law analysis of adiabatic combustion of ethylene


Consider the steady-flow adiabatic combustion of gaseous ethylene (C2H4) with
stoichiometric air, both at 25oC, and assume complete combustion. If T0=25oC, determine
(a) the entropy change and (b) the reversible work, both in kJ/kmol of fuel.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 207

This example is adopted from Wark [2], Example 10-5.

Solution
Figure E.1 shows the general Excel sheet that performs first-law analysis of the
combustion process with the data part modified to account for the present case in which
the fuel is ethylene (C2H4), γ =0%, Tin_air = 298K, and Pin_air =101 kPa. Since the
combustion process is adiabatic, Goal Seek was used to determine the adiabatic flame
temperature following the procedure of Example 8.5. The value determined by Goal
Seek, which is 2606.02K, is higher than that given by Wark [2], which is 2570K, but the
error is 1.4%.

Figure E.1. 1st law analysis of the adiabatic combustion of methane gas

Figure E.2 shows the sheet that performs second-law analysis of the combustion process.
The analysis is placed in a separate sheet with its data part that shows the specified values
of the surroundings pressure (Pref) and temperature (Tref), which are 101 kPa and 298K,
respectively. Based on these data, the sheet determines the rate of entropy production
(σ_production). The value determined by the sheet, which is 1190.3 kJ/kmol of C 2H4
agrees well with that given by Wark [2], which is 1189.55 kJ/kmol of C2H4.

Figure E.2. 2nd law analyses of the adiabatic combustion of methane gas
208 Mohamed M. El-Awad

The extension required by the present example is the determination of the rate of exergy
destruction (Xdestroyed) and reversible work (Wrev) using Equation (E.3). Since there
is no useful work in this process, the reversible work, which is the same as the exergy
destruction. The reversible work is determined as 354,713.6 kJ/kmol of C2H4, which is
close to the value given by Wark [2] of 354,660 kJ/kmol of C2H4.

Example E.2. Second-law analysis of isothermal combustion of methane


Methane (CH4) gas enters a steady-flow combustion chamber at 25°C and 1 atm and is
burned with 50 percent excess air, which also enters at 25°C and 1 atm. After combustion,
the products are allowed to cool to 25°C. Assuming complete combustion, determine (a)
the heat transfer per kmol of CH4, (b) the entropy generation, and (c) the reversible work
and exergy destruction. Assume that T0 = 298 K and the products leave the combustion
chamber at 1 atm pressure.

This example is adopted from Cengel and Boles [1], Example 15.11.

Solution
Figure E.3 shows the modified Excel sheet for this example in which T_ex is entered as
298K. Note is that for gaseous methane the enthalpy change (ΔhR_fuel) is automatically
adjusted to zero by the sheet. The value determined for the heat transfer is 871,617
kJ/kmol of fuel. This agrees well with the value obtained by Cengel and Boles [1], which
is 871,400 kJ/kmol of fuel.

Figure E.3. 1st law analysis of the isothermal combustion of methane gas

Figure E.4 shows the sheet that performs the second-law analysis of the combustion
process. The sheet calculates the entropy generation as 2,745.02 kJ/kmol.K. The value
obtained by Cengel and Boles [1] is 2,746.0 kJ/kmol.K. The exergy destruction, which
equals the reversible is calculated as 818,014.6 kJ/kmol of methane. This also agrees with
value given by Cengel and Boles [1], which is 818,000.0 kJ/kmol of methane.

For the special case considered in this example in which the reactants, the products, and
the surroundings are all at 25°C and the partial pressure Pi = 1 atm for each component
of the reactants and the products, Equation (E.3) reduces to [3]:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 209

W rev  N
R
~0
R g fR  N
P
~0
P g fP (E.8)

Where, g~ 0f is the Gibbs function of formation. Using Thermax functions that return values
of the Gibbs function of formation for the reactants and products, Figure E.5 shows that
the sheet determines the reversible work as 814,251.3 kJ/kmol of CH4, which is only
marginally different from the value obtained above by using Equation (E.3).

Figure E.4. 2nd law analyses of the isothermal combustion of methane gas
.

Figure E.5. Determining the reversible work from the difference in Gibbs function of
formation

The two previous examples applied the second law to analyse two important special cases
of adiabatic and isothermal combustion at atmospheric pressure. The following example
demonstrates the application of second-law analysis to the more general case in which
the process is neither adiabatic nor isothermal and the pressure is not atmospheric.
210 Mohamed M. El-Awad

Example E.3. Second-law analysis of a more general combustion process


1 kmole of butane enters a steady state steady flow reactor at 298 K and 250 kPa with
50% excess air. Combustion is assumed to be complete, and the products leave the reactor
at 1000 K and 250 kPa. Determine the heat transfer, reactant and product entropies, the
absolute availability of the reactants and products, the entropy change between the exit
and inlet, the entropy generation, the optimum work and the irreversibility.

This example is adopted from Annamalai and Puri. [4], Example 2 in Chapter 13.

Solution
Figures E.6 and E.7 show the sheets that perform first and second-law analyses of the
combustion process, respectively. Note that pressure (Pin_air) has been set to 250 kPa,
while the products exit temperature (T_ex) has been set to 1000K.

Figure E.6. First-law analysis of the combustion of butane

Figure E.7. Second-law analysis of the combustion of butane

Since butane is one of the fuels supported by Thermax, no special changes are required
to the calculations part. Figure E.6 shows that the heat transfer from the combustion
process is –1,531,670 kJ per kmole of butane. This compares well with the value given
by Annamalai and Puri. [4], which is –1,533,044 kJ per kmole of butane. The second-
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 211

law analysis sheet shows that 7,300.3 kJ/K of entropy is generated and 2,175,479 kJ of
reversible work is available per kmole of butane. The corresponding values given by
Annamalai and Puri. [4] are 7,294 kJ/K and 2,175,291 kJ, respectively. Since no actual
work has been produced, the irreversibility is equal to the reversible work, i.e., 2,175,479
kJ per kmol of butane.

References
[1] Y. A. Cengel and M.A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007
[2] K. Wark, JR, Advanced Thermodynamics for Engineers, McGraw Hill, 1995
[3] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics,
5th edition, John Wiley & Sons, Inc, 2006.
[4] K. Annamalai and I.K. Puri. Advanced Thermodynamics Engineering, CRC 2002
212 Mohamed M. El-Awad

Appendix F: Additional exercises and projects

This appendix provides additional exercises and projects related to thermodynamic


analyses of energy-utilisation systems. The majority of the cases considered in the
appendix are adapted from standard textbooks and published studies for the purpose of
getting more information about the basic principles involved and validating the analytical
models.

1. A 100-MWe steam power plant is to be designed on the basis of a reheat–regenerative


Rankine cycle with one open feedwater heater, one closed feedwater heater, and one
reheater as shown in Figure F.1. Steam enters the turbine at 15 MPa and 600 oC and
is condensed in the condenser at a pressure of 10 kPa. Some steam is extracted from
the turbine at P10 for the closed feedwater heater, and the remaining steam is reheated
at the same pressure to 600oC. 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 P12.

HP LP
Turbine Turbine

Steam 10 13
z
Generator 1-y-z
y 1-y 11
12
8 5
4 Condenser
7 2
6 3 1

Pump III Pump II Pump I


Figure F.1. A reheat–regenerative Rankine cycle with open and closed feedwater
heaters

(a) Develop an Excel sheet for thermodynamic analysis of the ideal cycle and verify
your sheet by comparing the values obtained for the fractions of steam extracted
from the turbine and the thermal efficiency of the cycle with the relevant values
given by Cengel and Boles [1] in Example 10.6 for P10 = 4 MPa and P12 = 0.5
MPa.
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 213

(b) Extend your Excel sheet to model the actual cycle by taking suitable values for
the isentropic efficiencies of the turbines and pumps.
(c) Use Solver to determine the optimum values of the two extraction pressures P10
and P12 that maximise the total annual revenue.
2. The use of closed feed-water heaters in a regenerative Rankine cycle requires a feed-
water pump to be added per each closed FWH (Refer to Figure 4.13). The added
pumps increase both the initial and running costs of the power plant. An alternative
arrangement that saves these additional costs is shown in Figure F.2.

Figure F.2

According to this arrangement, instead of pumping the feed-water to the next upper
level of pressure, the steam is throttled down to the feed-water heater with a lower
pressure. The trap allows only the saturated liquid to pass to the open feed-water
heater. Therefore, the vapour part has to be condensed first by losing its latent heat
of vaporisation. Develop an Excel sheet to analyse this cycle under the conditions
similar to those of Example 4.4 and compare the two cycles.
3. A vapour-compression refrigeration system includes a counter-flow double-pipe heat
exchanger as shown in Figure F.3. Ammonia leaves the evaporator as saturated
vapour at 1.0 bar and is heated in the heat-exchanger at constant pressure to 5oC
before entering the compressor. Following isentropic compression to 18 bar, the
214 Mohamed M. El-Awad

refrigerant passes through the condenser, exiting at 40oC, 18 bar. The liquid then
passes through the heat exchanger, entering the expansion valve at 18 bar.

3
Condenser

5
Compressor
2

Heat exchanger

Expansion
Valve 1

Evaporator
7

Figure F.3 Schematic of a VCR system with a counter-flow heat exchanger


(adapted from Moran and Shapiro [2], Problem 10.19)

If the mass flow rate of refrigerant is 12 kg/min, use Excel to determine:

a. the exit temperature of the subcooled refrigerant at state 6


b. the refrigeration capacity, in tons of refrigeration
c. the compressor power input, in kW
d. the coefficient of performance

Discuss possible advantages and disadvantages of this arrangement and study the
effect of the heat-exchanger effectiveness on the cycle’s performance by increasing
and decreasing the value of T1.
4. The ideal vapour-compression refrigeration cycle considered in Example 5.1 with
R-134a is modified to include a counter-flow double-pipe heat exchanger, as shown
in Figure F.3. The refrigerant leaves the evaporator as saturated vapour at -10.0oC
and is heated at constant pressure to state 2 before entering the compressor.
Following isentropic compression to P2 = 9 bar, the refrigerant passes through the
condenser, exiting at saturated liquid. The liquid then passes through the heat
exchanger where it is cooled to state 6 before entering the expansion valve at the
same pressure. The mass flow rate of refrigerant is 0.08 kg/s. Determine:
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 215

(a) the refrigeration capacity, in tons of refrigeration.


(b) the compressor power input, in kW.
(c) the coefficient of performance.

5. It is required to compare the performance of refrigerant R22 and refrigerant R134a


for heat-pump applications between an evaporator temperature of 10oC and a
condenser temperature of 70oC. Four options are possible:

a. A simple cycle with Refrigerant R22 alone


b. A simple cycle with Refrigerant R134a alone
c. A cascade system with R22 in the low-temperature cycle and R134a in the
high-temperature cycle
d. A cascade system with R134a in the low-temperature cycle and R22 in the
high-temperature cycle

Develop suitable Excel sheets to compare the compressor power (kW), heating
capacity (kW), and coefficient of performance of the above options. For options (c)
and (d), try various intermediate temperatures, e,g. 20, 30, 40oC
6. Methane, CH4, is burned with dry air. The fuel enters a combustion chamber at 400
K and 1 atm, where it is mixed with air entering at 500 K and 1 atm. The products
of combustion exit at 1800 K and 1 atm. The molar analysis of the products on a dry
basis is CO2, 9.7%; CO, 0.5%; O2, 2.95%; and N2, 86.85%.

(a) Show that the reaction equation is:

CH4 + 2.265 O2 + 11.515 N2 → 0.951 CO2+ 0.049 CO + 0.289 O2 + 11.515 N2


+ 2 H2O

(b) Determine the dew point temperature of the products, in oC, if the mixture
were cooled at 1 atm.
(c) For operation at steady state, determine the rate of heat transfer from the
combustion chamber in kJ per kmol of fuel. The average value for the
specific heat of methane between 298 and 400 K is 38 kJ/kmol K.

7. Liquid octane at 25oC, 1 atm enters a well-insulated reactor and reacts with air
entering at the same temperature (298K) and pressure (1 atm). For steady-state
operation and negligible effects of kinetic and potential energy, determine the
adiabatic flame temperature of the combustion products for complete combustion
with (a) the theoretical amount of air, (b) 400% theoretical air.
8. Use the general Excel sheet developed in Chapter 8 to study the effect of air
humidity on the heat release from the combustion process.
9. Use the general Excel sheet developed in Chapter 8 to study the effect of air
humidity on the adiabatic flame temperature of the combustion process
216 Mohamed M. El-Awad

10. Study the effect of excess air on the adiabatic flame temperature (Taf) of different
fuels by determining the values of Taf for isooctane (C8H18), propane (C3H8), and
methane (CH4) at γ = 0%, 10%, 20%, 30%, 40%, and 50%. Use a macro and plot
the results on an Excel chart.
11. Consider the simple vapour-compression refrigeration (VCR) system shown in
Figure F.4. The refrigerant can be assumed to leave the evaporator as saturated
vapour and the condenser as saturated liquid.

Qcond
Taci Taco

3 Condenser 2

Wcomp
Expansion
Compressor C
valve

Evaporator 1
4

Taeo Taei
Qevap

Figure F.4. Schematic diagram of a simple VCR system

It is required to optimise the total lifetime cost of the system for a specified value
of the cooling capacity Q . In optimisation analyses of the system, the total initial
e
cost of the system is weighed against the operating cost over its lifetime. While the
initial cost is paid once when the systems is being acquired, the operating costs
spreads throughout the system’s lifetime. For the system under consideration the
objective function for optimisation is that of the total cost (CT) given by:

CT  IC  PWOC (F.1)

Where IC is the initial cost that includes the costs of the compressor, evaporator,
and condenser, and PWOC is the present worth value of the annual operating cost.
The initial costs of the evaporator (ICE) and condenser (ICC), are given by:

IC evap  Ae  Cuae (F.2)


Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 217

IC cond  Ac  C uac (F.3)

Where Ae and Cuae are the area and cost per unit area for the evaporator and Ac and
Cuac are their counterparts for the condenser. The areas of the evaporator and
condenser depend on their rates of heat transfer, Q and Q , and their overall heat-
e c
transfer coefficients, Ue and Uc.

The initial cost of the compressor (ICComp) is given by:

C comp
IC comp  (F.4)
1   c 2
Where Ccomp is a unified reference cost of the compressor (the lowest possible cost
of a compressor) and ηc is the isentropic efficiency of the compressor.

The operating cost (OC), which is mainly the cost of electricity, can be obtained
from;

OC  W comp    C elec (F.5)

Where W comp is the compressor’s power in kW, τ is the annual operating time in
hours, and Celec the cost of electricity per kW-hr.

The present worth value of the annualised cost of electricity (PWOC) can be
obtained from:

PWOC 

OC  1  MARR   1
N
 (F.6)
MARR  1  MARR 
N

Where MARR is the minimum attractive rate of return for investment and N is the
life of the system in years.

Using Thermax or any other suitable property add-in, develop and Excel sheet that
can be used to optimise the system for an application of your choice using refrigerant
R134a, R22, R410A, or R717. Use your Excel sheet to optimise the VCR system for
a cooling capacity Q = 10 kW, lifetime n = 25 years, and MARR = 0.20 based on
e
the following data:
218 Mohamed M. El-Awad

Variable Name Value Units


Ccomp 25 $
Celec 0.10 $/kW-hr
Cuac 200 $/m2
Cuae 350 $/m2
Time 3000 kW-hr/yr
o
Taci 30 C
o
Taei 20 C
Uc 0.20 kW/m2-K
Ue 0.30 kW/m2-K

This case is based on a paper contributed by Dr. David C. Zietlow [3].

12. Figure F.5 shows a schematic diagram of a combined power-generation system that
consists of a compound diesel engine (compressor C1, intercooler IC1, diesel engine
DE, and turbine T1) and an intercooled regenerative gas turbine (compressor C2,
intercooler IC2, compressor C3, regenerator RG, combustion chamber CC, and
turbine T2). Before discharged to the atmosphere, the exhaust gases from the two
turbines are mixed and passed through the regenerator to reheat the high-pressure
air going to the combustion chamber of the gas turbine. Figure F.6 shows the T-s
diagram of the combined system.

Figure F.5. Schematic diagram of the combined diesel-engine gas-turbine system


(adopted from El-Awad [4])
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 219

Figure F.6. T-s diagram of the combined diesel-engine gas-turbine system (adopted
from El-Awad [4])

Develop an Excel sheet to analyse the performance of the combined cycle with specified
values of the diesel-engine compression ratio (CR) and the turbine inlet temperature (TIT)
of 1000K and 20, respectively. Treat the post-combustion products of the diesel engine
and gas turbine as air and base your analysis on the following condition:

- Ambient temperature and the pressure are 300 K and 0.1 MPa, respectively.
- The diesel engine's cut-off ratio (COR) is 2.
- The regenerator effectiveness (ε) is 80%.
- The mass flow rates in both the diesel engine (m1) and the gas turbine (m2) are
1 kg/s, i.e. m2= m1=1 kg/s.

Use Excel's Solver to determine the pressure ratios of the three compressors that
maximise the cycle's thermal efficiency and net work-output.

References
[1] Y. A. Cengel and M.A. Boles. Thermodynamics an Engineering Approach,
McGraw-Hill, 7th Edition, 2007.
[2] M.J. Moran and H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 5th
edition, John Wiley & Sons, Inc, 2006.
[3] D. C. Zietlow, Optimization of vapor compression cycles, 2014 ASEE Annual
Conference & Exposition, 24.958. 1-24.958. 23
[4] M.M. El-Awad, Energy and Exergy Analysis of a combined diesel-engine gas-
turbine system for distributed power generation. Int. J. of Thermal & Environmental
Engineering, Vol. 5, No. 1, pp. 31-39 (February 2013).
220 Mohamed M. El-Awad

Subject Index

acentric factor, 180 macros, 9, 154, 155, 156


adiabatic flame temperature, 9, 138, multi-stage compression, 84, 102,
151, 152, 154, 155, 156, 191, 184
192, 195, 203, 204 optimisation analyses
adiabatic mixing, 116, 131, 133 Solver, 4, 7, 9, 204
air-conditioning, 6 Optimisation analyses, 9
ammonia-water, 10, 20, 21, 31, 32, Otto cycle, 7, 10, 36, 44, 45, 46, 47,
33, 158, 160, 161, 162, 169, 170 48, 49, 50
Brayton, 32, 36, 37, 38, 40, 41, 42, R134a, 8, 9, 86, 88, 89, 91, 99, 205
43, 71, 72, 73, 74, 79, 81 R22, 9, 89, 99, 205
Brayton cycle, 2, 36, 37, 38, 40, 41, Rankine cycle, 2, 9, 10, 53, 54, 56,
42, 43, 72, 73, 74 57, 58, 59, 60, 61, 62, 64, 67, 68,
cascade refrigeration systems, 114 69, 71, 72, 74, 75, 79, 80, 81,
combined cycle, 72, 75, 76, 77, 78, 100, 200, 201
79, 207 Raphson, 8, 9
combustion analyses, 10, 14, 22, regenerative gas turbine, 51, 52, 206
137, 138, 139, 143, 144, 145, regenerative Rankine cycle, 9, 54,
146, 154, 156, 186, 191, 193, 194 79
cooling tower, 125, 126 Saturated steam, 56
dew-point temperature, 117, 138, single-effect VAR system, 10
140, 141, 142, 144, 146, 156 Soave, 8, 180
Diesel cycle, 7, 44, 51, 52 stoichiometric air-fuel ratio, 22, 140
double-effect VAR system, 166 stoichiometric combustion, 139,
double-pipe, 201, 202 147, 153
evaporative cooler, 116, 122, 123 Superheated steam, 60
Excel formula, 26, 57, 61 Thermax1, 10, 23, 24, 25, 26, 30,
Exergy, 47, 48, 49, 78, 79, 193, 207 31, 33, 38, 42, 55, 56, 57, 60, 75,
feed-water, 5, 6, 72, 75 79, 114
formula bar, 57 vapour-absorption refrigeration, 10,
heat-exchanger, 40, 202 14, 16, 30, 157, 158, 169, 170
hydrocarbon fuel, 138, 139, 140, vapour-compression refrigeration,
141, 144, 156 10, 84, 99, 102, 158
ideal gas Cycle, 84, 85, 86, 87, 88, 99, 201,
law, 8, 11, 16, 27, 29, 177 202, 204
Iterative solutions VBA, 32, 79
Goal Seek, 9 wet cooling tower, 116, 125, 126
Lithium-bromide-water, 10 X-Steam, 39, 75
Thermodynamic Analyses of Energy-Utilisation Systems Using Excel 221

View publication stats

You might also like