You are on page 1of 98

ENERGY AND EXERGY ANALYSIS OF WASTE

HEAT RECOVERY SYSTEMS USING


ORGANIC RANKINE CYCLE

A thesis submitted to
Department of Mechanical Engineering,
Bangladesh University of Engineering and Technology (BUET)
Dhaka

by

Anup Saha

IN PARTIAL FULFILMENT OF THE REQUIREMENTS


FOR
THE DEGREE OF MASTER OF SCIENCE
IN
MECHANICAL ENGINEERING

August, 2016

i
Certificate of Approval

The thesis titled, ENERGY AND EXERGY ANALYSIS OF WASTE HEAT RECOVERY
SYSTEMS USING ORGANIC RANKINE CYCLE, submitted by Anup Saha, Roll No:
1014102012 P, Session: October/2014, has been accepted as satisfactory in partial fulfillment
of the requirements for the degree of MASTER OF SCIENCE IN MECHANICAL
ENGINEERING on 20 August, 2016.

ii
Candidates Declaration

I, hereby, declare that this thesis or any part of it has not been submitted elsewhere for the
award of any degree of diploma. And also declare that all information in this document has
been obtained and presented in accordance with academic rules and ethical conduct. I also
declare that, as required by these rules and conduct, I have fully cited and referenced all
material and results that are not original to this work.

Anup Saha

iii
Acknowledgements

I would like to express my sincere appreciation to my supervisor, Professor Dr. Md. Zahurul
Haq whose expertise, understanding, generous guidance and support made it possible for me
to work on a topic that was of great interest to me. For his unwavering support, I am truly
grateful. He always knew where to look for the answers to obstacles while leading me to the
right source, theory and perspective.

A debt of gratitude is owed to my parents and brother, without their love and support, I would
not have come this far.

I would like to thank Guido van Rossum and all others associated with the development of
Python programming language.

Finally, I wish to thank the Almighty for making me who I am.

iv
Abstract
Exergy analysis of thermodynamic systems along with the energy analysis is growing rapidly.
Exergy quantifies the potential of available energy to do useful work. Exergy analysis is
fruitful to identify processes where exergy destruction and losses occur and acts as a useful
supplement to the energy analysis. Utilization of waste heat has attracted enormous attention
in recent years due to environmental concerns. Organic Rankine Cycle (ORC) is a promising
technology for conversion of low grade heat (at low temperature) into useful work. In the
present study, detailed energy and exergy analysis of a bottoming ORC, utilising the waste
heat, is performed. Suitable refrigerants (R141b, R124, R21, R1233zd(E), R1234ze(Z) and
R245fa) having low Global Warming Potential (GWP) values and compatible with the thermal
condition of the heat source are selected as the working fluids for the analysis. Effects of
evaporator pressure on net work output, power consumption, mass flow rate, thermal and
exergy efficiency, irreversibility of individual components as well as of the overall system is
investigated. Both thermal and exergy efficiencies of the ORC increase with the evaporator
pressure. Among the refrigerants considered, R141b results in the highest efficiency due to its
highest pressure ratio, whereas R124 results in the lowest (pressure ratio, thermal efficiency
and exergy efficiency for R141b are 45.1, 19.3% and 48.5% respectively at 60 bar evaporator
pressure, whereas for R124, the corresponding values are 10.1, 12.0% and 30.2% respec-
tively). Energy balance indicates that major portion of the input energy is rejected to
condenser cooling water, although exergy balance suggests that its exergy content is very low
(using R245fa at 10 bar evaporator pressure, heat rejection is 89% of total energy input but its
exergy value is only 7% of total input exergy). However, heat rejection to condenser cooling
water decreases with increasing evaporator pressure (the heat rejection value decreases by
9.1% when evaporator pressure is increased from 10 to 65 bar using R245fa). A major part of
the total input exergy is destructed due to the irreversibility in various components of the sys-
tem. Among the components, the highest exergy destruction occurs in the condenser because
of higher temperature difference of heat exchanging fluids (using R245fa at 10 bar evaporator
pressure, 70% of the total exergy destruction occurs in the condenser). Consequently, con-
denser has the lowest exergy efficiency among other components which is responsible for low
overall efficiency of the system. The results also demonstrate that, commonly used refrigerant
R245fa can be potentially replaced by low GWP refrigerants R1233zd(E) and R1234ze(Z)
with slight variations of the performance of ORC.
v
List of Symbols

ORC Organic Rankine Cycle

Exw Exergy associated with work transfer

ExQ Exergy associated with heat transfer

Exth Thermo-mechanical exergy

Exch Chemical exergy

Extotal Total exergy

Ed Exergy destruction

g0 Standard value of Gibbs function of the reaction

P0 Atmospheric Pressure

T0 Ambient temperature

ef Specific flow exergy

Wsurr Surrounding work

h Specific enthalpy of the stream at the given state

s Specific entropy of the stream at the given state

u Specific internal energy of the stream at the given state

v Specific volume of the stream at the given state

Chemical potential of the substance

Cp Specific heat at constant pressure

R Gas constant

n Number of moles

I First law efficiency

II Second law efficiency

m Mass flow rate of the fluid stream

vi
W Rate of work

Q Rate of heat transfer

x Quality

OMTS Octamethyl-Trisiloxane

EOS Equation of State

Subscripts:

r Refrigerant

i Inlet condition

e Exit condition

0 Restricted equilibrium with the environment

00 Unrestricted equilibrium with the environment

cv Control volume

hs Heat source

eva Evaporator

p Pump

t Turbine

con Condenser

vii
Table of Contents
Certificate of Approval.............................................................................................ii

Candidates Declaration......................................................................................... iii

Acknowledgements.................................................................................................. iv

Abstract .................................................................................................................... v

List of Symbols ......................................................................................................... v

Table of Contents...................................................................................................viii

List of Figures.......................................................................................................... xi

List of Tables .........................................................................................................xiii

1 Introduction..................................................................................................... 1

1.1 Energy and Exergy ..................................................................................... 1


1.2 Organic Rankine Cycle (ORC) and Waste Heat Recovery .......................... 2
1.3 Objective of the Thesis ............................................................................... 4
1.4 Outline of the Thesis .................................................................................. 5
2 Literature Review ........................................................................................... 7

2.1 Thermodynamic Analysis of Energy Conversion Systems .......................... 7


2.2 Evolution of Organic Rankine Cycle (ORC) ............................................... 9
2.3 Key properties of potential ORC fluids..................................................... 10
2.4 Working fluid selection for ORC .............................................................. 12
2.5 Bottoming ORC ....................................................................................... 15
3 Exergy ........................................................................................................... 17

3.1 Exergy and its Physical Meaning .............................................................. 17


3.2 Components of Exergy of a Thermodynamic System ............................... 19
3.2.1 Exergy associated with work transfer (Exw) ......................................... 19
3.2.2 Exergy associated with heat transfer (ExQ) ........................................... 20
3.2.3 Thermo-mechanical exergy (Exth) ........................................................ 20
3.2.4 Chemical Exergy (Exch ) ...................................................................... 22
viii
3.3 Exergy Balance of Thermodynamic Systems ............................................ 23
3.3.1 Exergy Destruction (Irreversibility) ..................................................... 24
3.3.2 Second Law Efficiency ........................................................................ 24

4 Thermodynamic Modelling ......................................................................... 26

4.1 Thermodynamic Modelling of ORC system ............................................. 26


4.2 Mathematical Model ................................................................................ 28
4.2.1 Energy and Exergy analysis of the cycle components ........................... 28
4.2.1.1 Energy Analysis ............................................................................ 28
4.2.1.2 Exergy Analysis ............................................................................ 28
4.2.2 Thermodynamic Relationships in the ORC .......................................... 30
4.2.2.1 Energy balance equations for ORC ............................................... 30
4.2.2.2 Exergy balance equations for ORC ............................................... 32
4.2.2.3 Overall System Analysis ............................................................... 33
5 Simulation and Results ................................................................................. 35

5.1 Validation of the Model developed ........................................................... 35


5.1.1 Validation of Topping Cycle (GT Cycle) .............................................. 35
5.1.2 Validation of Bottoming ORC model ................................................... 37
5.2 Modelling Assumptions ............................................................................ 40
5.3 Screening of Potential working fluids ....................................................... 41
5.4 Drawback of water over organic fluids ..................................................... 42
5.5 Parametric Effect on the performance of ORC .......................................... 44
5.5.1 Energy and Exergy Balance of ORC .................................................... 45
5.5.2 Exergetic efficiency of components at 10 bar ....................................... 47
5.6 ORC performance analyses at different evaporator pressures for R245fa .. 48
5.6.1 Net Power Output and Power Consumption for R245fa ....................... 48
5.6.2 Mass flow rate for R245fa ................................................................... 49
5.6.3 Overall Efficiency of ORC for R245fa ................................................. 50
5.6.4 Irreversibility in components for R245fa.............................................. 51
5.6.5 Exergy Efficiency variation of ORC components for R245fa ............... 53
5.7 Performance of ORC for various refrigerants ........................................... 55
5.7.1 Net Power Output ................................................................................ 56
5.7.2 Thermal Efficiency .............................................................................. 57
5.7.3 Heat rejection and flow rate of coolant at Condenser ........................... 58
ix
5.8 Cycle Irreversibility Analysis ................................................................... 60
5.8.1 Second Law / Exergy Efficiency of Overall System ............................. 60
5.8.2 Exergy carried out by condenser cooling water .................................... 61
5.8.3 Exergy Destruction of ORC components ............................................. 61
5.8.4 Exergy Efficiencies of ORC components ............................................. 65
6 Conclusions.................................................................................................... 68

Appendix A ............................................................................................................. 70

Appendix B ............................................................................................................. 78

References ............................................................................................................... 81

x
List of Figures
Figure 1.1 Organic Rankine Cycle combined with GT cycle .............................................. 3
Figure 2.1 Cumulative installed capacity of ORC plants [30] ............................................. 9
Figure 2.2 Dry, Isentropic and Wet working fluids [43] .................................................... 13
Figure 3.1 Exergy Transfer with Heat .............................................................................. 20
Figure 3.2 Components of Exergy [68] ............................................................................ 21
Figure 4.1 Schematic Illustration of the ORC ................................................................... 30
Figure 4.2 Schematic of ORC evaporator ......................................................................... 31
Figure 4.3 Schematic of ORC pump................................................................................. 31
Figure 4.4 Schematic of ORC turbine .............................................................................. 31
Figure 4.5 Schematic of ORC condenser .......................................................................... 32
Figure 5.1 Energy balance Comparison with System A of Literature [74]....................... 36
Figure 5.2 Comparison with refrigerant flow rates of O. Kaska [26]................................ 38
Figure 5.3 Comparison with heat available of O. Kaska [26]........................................... 38
Figure 5.4 Comparison with turbine power of O. Kaska [26]........................................... 39
Figure 5.5 Comparison with pump power of O. Kaska [26]............................................. 39
Figure 5.6 Comparison of wet fluid and dry fluid [79] ..................................................... 42
Figure 5.7 T-s diagram of water ....................................................................................... 43
Figure 5.8 T-s diagram of several organic fluids ............................................................... 43
Figure 5.9 Component-wise energy consumption for R245fa at 10 bar evaporator pressure
.45
Figure 5.10 Energy Balance for the system for R245fa at 10 bar evaporator pressure ..... 46
Figure 5.11 Exergy Balance for the system for R245fa at 10 bar evaporator pressure...... 46
Figure 5.12 Exergy Efficiency of components for R245fa at 10 bar evaporator pressure . 47
Figure 5.13 Work Output and Consumption with evaporator pressure for R245fa ........... 49
Figure 5.14 Refrigerant and Coolant flow rate with evaporator pressure for R245fa ....... 50
Figure 5.15 Exit temperature variation with evaporator pressure for R245fa ................... 50
Figure 5.16 Variation of Overall Efficiencies of ORC for R245fa ................................... 51
Figure 5.17 Distribution of exergy destruction with evaporator pressure for R245fa ....... 52
Figure 5.18 Exergy Efficiency of the ORC components for refrigerant R245fa............... 54
Figure 5.19 Variation of performance parameters with evaporator pressure .................... 55
Figure 5.20 Net work output variation for selected refrigerants ...................................... 56
xi
Figure 5.21 Pump consumption variation for selected refrigerants .................................. 57
Figure 5.22 Thermal efficiency variation for selected refrigerants .................................. 58
Figure 5.23 Condenser rejection variation for selected refrigerants ................................. 59
Figure 5.24 Expander outlet temperature variation for selected refrigerants .................... 59
Figure 5.25 Coolant flow rate variation for selected refrigerants ..................................... 60
Figure 5.26 Exergy efficiency variation of selected refrigerants ...................................... 61
Figure 5.27 Exergy carried out by condenser cooling water for selected refrigerants ...... 62
Figure 5.28 Evaporator exergy destruction for selected refrigerants ................................ 63
Figure 5.29 Condenser exergy destruction for selected refrigerants ................................ 63
Figure 5.30 Pump exergy destruction for selected refrigerants ........................................ 64
Figure 5.31 Turbine exergy destruction for selected refrigerants ..................................... 64
Figure 5.32 Evaporator exergy efficiency variation for selected refrigerants ................... 65
Figure 5.33 Turbine exergy efficiency variation for selected refrigerants ........................ 66
Figure 5.34 Condenser exergy efficiency variation for selected refrigerants .................... 66

xii
List of Tables

Table 3.1 Reference Substances [1] ................................................................................ 18


Table 5.1 Energy Balance (MW) Comparison with System A of [74] .......................... 36
Table 5.2 Exergy Balance (MW) Comparison with System A of [74] ........................... 36
Table 5.3 Comparison between the results of [75] and Present Study .............................. 37
Table 5.4 Properties of Screened Organic Fluids ............................................................. 41
Table 5.5 Representative Energy and Exergy performance data for R245fa at 10 bar ...... 44
Table 5.6 ORC parameters at different evaporator pressures for R245fa ......................... 48
Table 5.7 Exergy Destruction of components for R245fa ................................................ 52
Table 5.8 Saturation pressure variation of refrigerants at condensing temperature........... 55
Table A.1 ORC parameters at different evaporator pressure for R141b ............................ 70
Table A.2 Exergy Destruction of ORC components for R141b ........................................ 70
Table A.3 Exergy Efficiency variation of ORC components for R141b............................ 71
Table A.4 ORC parameters at different evaporator pressure for R124 .............................. 72
Table A.5 Exergy Destruction of ORC components for R124 .......................................... 72
Table A.6 Exergy Efficiency variation of ORC components for R124 ............................. 73
Table A.7 ORC parameters at different evaporator pressure for R21 ................................ 73
Table A.8 Exergy Destruction of ORC components for R21 ............................................ 74
Table A.9 Exergy Efficiency variation of ORC components for R21 ............................... 74
Table A.10 ORC parameters at different evaporator pressure for R1233zd(E) ................... 75
Table A.11 Exergy Destruction of ORC components for R1233zd(E)................................ 75
Table A.12 Exergy Efficiency variation of ORC components for R1233zd(E) ................... 76
Table A.13 ORC parameters at different evaporator pressure for R1234ze(Z).................... 76
Table A.14 Exergy Destruction of ORC components for R1234ze(Z) ................................ 77
Table A.15 Exergy Efficiency variation of ORC components for R1234ze(Z) ................... 77

xiii
1 Introduction

1.1 Energy and Exergy


Energy is conserved in every device or process. The concept of efficiency in the traditional
sense is based on the first law of thermodynamics. The drawback of the first law is that it
accounts for only the quantity of energy. The quality of energy in the thermodynamic sense is
synonymous to its ability to be transformed to useful work, as stipulated by the second law of
thermodynamics, which bestows a much greater glory on ordered forms of energy (e.g. work)
than disordered form of energy (e.g. heat) [1]. So, when comparing two different kinds of
energy transformations, some clarification is needed with the definition of efficiency.
According to the second law of thermodynamics, energy has quality as well as quantity. The
second law provides an additional means of evaluating and comparing processes and systems
meaningfully by introducing exergy [2].

Exergy is defined as the maximum theoretical work that can be obtained from a combined
system (combination of a system and its reference environment) when the system comes into
complete thermodynamic equilibrium (as thermally, mechanically and chemically) with the
environment without violating any laws of thermodynamics [3]. The maximum available
work from a system emerges as the sum of two contributions: thermo-mechanical exergy and
chemical exergy. Thermo-mechanical exergy is defined as the maximum extractable work
from the combined system as the system comes into thermal and mechanical equilibria with
the environment, and its value is calculated with respect to restricted dead-state condition. At
the restricted dead-state conditions, no work potential exists between the system and the envi-
ronment due to temperature and pressure differences. In this state, the control mass is in
thermo-mechanical equilibrium with the environment, but not necessarily in chemical equi-
librium with it [3]. In principle, the difference between the compositions of the system at the
restricted dead-state and the environment, and the difference in species concentrations
between the system and environment can be used to obtain additional work to reach chemical
equilibrium through reaction and diffusion respectively. The maximum additional work
obtained in this way is called the chemical exergy [1].

Unlike energy, exergy (also known as availability) is not a conserved entity. Every irreversi-
ble process causes a loss in availability which is irrecoverable as dictated by the second law

1
of thermodynamics. Since no natural process is totally reversible, we always have to accept
some form of loss in availability no matter how hard we try. But acceptance of loss in avail-
ability should always be compensated by some economic justification. Non-existence of such
a justification indicates that the availability loss results only from an error in the art of engi-
neering. Thus presence of an availability loss invariably indicates the opportunity of a
thermodynamic improvement [4]. So exergy analysis is crucial to identify sites where exergy
destruction and losses occur and rank order them for significance [5].

1.2 Organic Rankine Cycle (ORC) and Waste Heat Recovery


In a typical developed country as much as 40% of the total fuel consumption is used
for industrial and domestic space heating and process heating. Of this, around one third is
wasted. Low grade heat has generally been discarded by industry and has become an
environmental concern because of thermal pollution. This wasted heat can be lost to the
atmosphere at all stages of a process, through inefficient generation, transmission, or during
final use of the energy. This has led to the search for technologies which not only reduce the
burden on non-renewable sources of energy but also take steps toward a cleaner envi-
ronment. Also, given the growing scarcity of primary energy resources, achieving increased
efficiency of energy conversion processes is one of the key challenges for optimising primary
energy use. From this perspective, low temperature waste heat from various processes is
becoming more and more attractive as a secondary energy source [6].

Waste heat can be recovered either directly or more commonly, indirectly. Direct heat
recovery is often the cheaper option, but its use is restricted by location and contamination
considerations. In indirect heat recovery, two fluid streams are separated by a heat transfer
surface. Devices that convert low grade heat to electricity and can be retro-fitted to existing
plants to increase their efficiency and contribute to their emission reductions are of
great interest. Used in this way, technologies that convert low grade heat to electricity
can be advantageous on two fronts. Firstly, by the improvement of the efficiency of cur-
rent technology and secondly, in application to sustainable energy sources that are, to date,
unexploited.

In recent years, the performance of gas turbine (GT) has been significantly improved [7].
However, the efficiency of conventional power plants, which are usually based on single
prime movers is usually less than 40% [8]. Most of the input energy is lost as exhaust waste
heat. To harness this significant waste energy, combined cycles are popular choice [9]. But, if
2
the exhaust temperature is low/medium, utilizing the conventional steam Rankine cycle (RC)
to recover waste energy is not the best choice due to dissatisfactory performance [10, 11].

In a typical GT plant, to produce an output of 1 MW, the exhaust from the GT contains ap-
proximately 1.6 MW heat. In conventional combined cycle power plant (CCPP),
approximately 0.6 MW of this exhaust waste heat can be recovered.

ORC or Organic Rankine cycle basically resembles the steam Rankine Cycle according to
working principles. In ORC, Water is replaced with a high molecular mass fluid with lower
degree of boiling temperature in comparison with water. Fluid characteristics make ORC fa-
vorable for applications of low temperature heat recovery.

Figure 1.1 Organic Rankine Cycle combined with GT cycle

The ORC functions in a similar way as the conventional steam Rankine Cycle. The principle
is simple. The organic fluid is pumped into a heat exchanger where its vaporized. The high
3
pressure vapour flows through an expander and outputs technical work due to the pressure
drop. The exhaust from the expander goes into the other heat exchanger where its condensed,
and then returns to the pump.

One approach which is found to be highly effective in addressing the above mentioned issues
is to make use of low grade heat to generate electric power in an Organic Rankine cycle
(ORC) system. For low to medium temperature heat sources, organic working fluids offer
advantages over water as the working medium, as used in conventional Rankine cycle
systems, by increasing the cycle efficiency, thereby enabling more power to be generated.
This has been shown to be particularly promising for decentralized combined heat and power
production.

The recovery of waste heat has a direct effect on the efficiency of the process. This results in
both reduced utility consumption and process costs. It also reduces the fuel consumption,
which leads to reduction in the flue gas produced. This permits equipment sizes of all flue gas
handling equipment such as fans, stacks, ducts, burners, etc. to be reduced in addition to re-
ducing atmospheric pollution [6].

One significant advantages of ORC and gas turbine is that they can compete with high tem-
perature exhaust gas turbine even if the temperature is lower, hence this will give the
opportunity to have gas turbines with lower turbine inlet temperature, less combustion tem-
perature and consequently less NOx production, manufacturing and operational cost. Gas
turbine will be cheaper since less resistant material and cooling methods will be necessary
[12].

ORC has also the advantage of using a number of refrigerants as working fluid depending on
the thermal conditions of heat source. Attending to environmental issues, analysing the ORC
with new environment-friendly refrigerants are being emphasized on recent researches [13,
14].

1.3 Objective of the Thesis


Present study is focused on detailed energetic and exergetic analysis of a bottoming ORC
running with several newly developed refrigerants as working fluid to achieve optimum en-
ergy/exergy efficiency utilizing of the waste energy/exergy associated with GT exhaust.

4
The objectives of this study can be enumerated as:

1. To develop a thermodynamic model of a bottoming ORC driven from GT


exhaust.
2. To carry out exergy analysis in conjunction with energy analysis of the ORC
model developed.
3. To select suitable organic working fluid candidates compatible with the
thermal condition of the heat source and also considering their impact on the
environment e.g., Global Warming Potential (GWP).
4. To figure out the second law efficiency / exergetic efficiency along with
energetic efficiency for the modified cycle.
5. To quantify exergy destruction of components as well as the overall cycle.

1.4 Outline of the Thesis


The present work takes a thermodynamic approach to analyse bottoming ORC using GT
exhaust gases. Since the exergy method is ideally suited for this kind of analysis, majority of
the post-processing and analyses are done using exergy method.

The most relevant literature on ORC evolution, ORC working fluid selection, thermodynamic
analyses of bottoming ORC have been reviewed in chapter 2 in chronological order.

Comprehensive exergy analyses of thermodynamic systems require calculation of both physi-


cal and chemical exergies. In chapter 3, all the required terminology and concepts for the
development of exergy balance equations are first discussed. Then, balance equations are
developed for both control mass and control volume systems. Special attention has been paid
to integrate chemical exergies of the contents of the system by ensuring calculation of both
reactive and diffusive components.

Detailed thermodynamic modelling and simulation procedures of the considered bottoming


ORC are discussed in chapter 4. Energy and exergy based analyses of the ORC are modelled
here from the governing thermodynamic equations.

Chapter 5 starts with the validation of the model with the available experimental data. Next,
suitable organic fluids are selected based on thermal condition of the heat source and
environmental impact. Selected fluids are tested one by one and variations of key perform-
ance parameters such as net power output, power consumption, mass flow rate, energy and
exergy efficiencies, exergy destruction of the system as well as components are analysed. In
5
the exergy analysis, component-wise exergy destruction/losses were evaluated to determine
the distribution of the exergy destruction in the system.

6
2 Literature Review

Fossil fuel resources are finite. The three major types of fossil fuel, namely, coal, oil, and
natural gas, were formed millions of years ago when dead plants and animals were trapped
under deposits and became buried underneath land. They cannot be replenished once used. As
reported by BP, the proved reserves of coal, oil, natural gas were 860,938 million tons,
1668.9 billion barrels and 6614.1 trillion cubic feet at the end of 2012. The corresponding
reserves/production (R/P) ratio was 109, 52.9 and 55.7 years [15]. Furthermore, generation
of electricity by burning fossil fuels puts a significant strain on the environment. Therefore it
is essential to increase conversion efficiencies to optimally exploit the potential of our re-
sources. This is the landscape where the Organic Rankine Cycle (ORC) comes into play.

2.1 Thermodynamic Analysis of Energy Conversion Systems


Energy in our world is found manifested in many forms, each with its own characteristics and
its quality. The historical acceptance of the quality of energy finds itself in performing
mechanical work. Hence, quality of energy today is synonymous with its capacity to cause
change or do useful work. Natural observation shows that particular forms of energy differ in
their ability to be transformed into other forms. Moreover, this ability is found to be depend-
ent on composition and state properties of both system and surrounding. It is then only
rational to have the quality index of energy acknowledges the transformability of a certain
form of energy with respect to the environment; a fact overlooked by the universal law of en-
ergy conservation. Exergy bridges this gap in the analysis of energy transformations by using
the second law of thermodynamics, which imposes certain restrictions in the direction and
amount of energy transformations. Exergy is a measure of the maximum derivable work out-
put (or, the minimum required work input) from a given thermodynamic process with
specified conditions for both system and surrounding [1].

Though the quest for qualitative energy index can be traced back to the work of Gibbs [16] as
available energy, the term exergy was first coined by Zoran Rant in 1953 from Greek root
words [17] . Exergy means that fraction of the total energy that can be extracted as work. The
concept of exergy has been greatly explored and brought up to the applied sciences over the
past 50 years.

7
Much of the earlier works on the developments in the applied level can be attributed to Szar-
gut and Kotas in the eighties [1, 18]. Specially the works on standard chemical exergy of the
elements and some compounds by Szargut et al. has to be mentioned as one of the seminal
works in this field [19]. On the other hand, there has been a parallel theoretical development
of the exergy concept from pure thermodynamic point of view by Haywood [20] and Suss-
man [21, 22].

It was not long before the concepts of second law analysis and exergy destruction started ap-
pearing in ORC and combined cycle. But in the recent past there was tremendous
contribution in the exergy based analysis for combined cycle. There are many researchers
such as Kotas [1] and Moran-Shapiro [5] who carried out the exergy analysis for the com-
bined cycles. They found out the exergy losses in every part.

CHP for geothermal resources of low and medium quality were compared by Florian and Di-
eter [23] to the option of power generation by a second law analysis. Second law efficiencies
are analyzed for a variety of parameters like temperature of geothermal water, supply tem-
perature of the heating system and demanded thermal power.

R. Long et al. [24] adopted the internal and external exergy efficiencies to analyze the impact
of working fluids on the performance of the organic Rankine cycle, and a simplified internal
exergy efficiency model is proposed to indicate this impact. The calculation results showed
that the thermo-physical properties of the working fluid have little impact on internal exergy
efficiency, but they do play an important role in determining external exergy efficiency.

Wang et al. [25] analysed a double organic rankine cycle for discontinuous waste heat recov-
ery. The influence of outlet temperature of heat source on the net power output, thermal
efficiency, power consumption, mass flow rate, expander outlet temperature, cycle irreversi-
bility and exergy efficiency at a given pinch point temperature difference (PPTD) were
analysed. Variations of energy and exergy efficiencies of the system with evapora-
tor/condenser pressures, superheating and subcooling were illustrated by O. kaska [26] . It
was observed from the analysis that, the energy and exergy efficiencies of the system are
10.2%; 48.5% and 8.8%; 42.2%, respectively, for two different actual cases. Exergy destruc-
tion of subcomponents was also quantified.

Wu et al. [27] proposed a model of evaporator in ORC, recovering waste heat, based on ex-
ergy recovery. The study contributed an original approach, that can be useful in the design of
evaporator for waste heat recovery based power generation. Karellas et al. [28] presented
8
waste heat recovery as a way to gain energy from the exhaust gases in a cement plant. Water
steam cycle and ORC has been analyzed for waste heat recovery. The energetic and exergetic
evaluation of the two waste heat recovery processes was presented and compared.

2.2 Evolution of Organic Rankine Cycle (ORC)


It is believed that idea of ORC as power generation system from low heat sources was first
presented in 1961 by Israeli solar engineers Harry Zvi Tabor and Lucien Bronicki [29].
Though it is difficult to clarify the exact time the concept of ORC came up, it is clear the
practical usage of the ORC is accelerated by the need to become less dependent on fossil fuel.

Since the oil-shock in 1970s, there have been significant developments in the ORC. The ad-
vantage of ORC in low temperature and power applications is a good match for heat sources
from the combustion of biomass, geothermal energy, and industrial exhaust, etc. The collec-
tion, transportation, and storage costs motivate decentralized power generation [30].

Hundreds of ORC plants have operated reliably for many years all over the world. Some in-
formation is presented. By the end of 2013, the installed capacity of ORC power plants has
climbed to approximately 1,700 MW. And the growth gets faster and faster, as shown in Fig-
ure 2.1[30].

Figure 2.1 Cumulative installed capacity of ORC plants [30]

ORC manufacturers have been present on the market since the beginning of the 1980s. They
provide ORC solutions in a broad range of power and temperature levels, as shown in. Note
that only manufacturers with several commercial references have been detained in this sur-
vey. At present Turboden and Ormat are two the largest manufacturers in the field. The
9
former has constructed more than 200 biomass ORCs and the latter has built up more than 30
geothermal ORCs. Triogen claims the most efficient ORC in the market with an efficiency of
>17 % (150 kW) and also the most cost-effective one measured in cost per kW installed ca-
pacity. And the payback period is between 25 years [31]. The ORC technology in power
range above 100 kW has reached a considerable degree of maturity.

2.3 Key properties of potential ORC fluids


Numerous research has been conducted regarding the properties of working fluids in the
ORC that should be considered for the design and selection of working fluids for ORC
processes. A non-exhaustive list of desired properties of ORC fluids gleaned from several re-
searches is described below [11, 32-37] :

Global warming potential (GWP):

The global warming potential is an index that determines the potential contribution of a
chemical substance to global warming. Hence, the refrigerant should have low environ-
mental impact and greenhouse warming potential.

Toxicity of working fluid:

All organic fluids are inevitably toxic. A working fluid with a low toxicity should be used to
protect the personnel from the threat of contamination in case of fluid leakage. Hence, the
determination of the toxicity of the designed working fluids is important for human
safety reasons.

Density:

This parameter is of key importance, especially for fluids showing a very low con-
densing pressure. The density of the working fluid must be high either in the liquid
or vapour phase. High liquid or vapour density results to increased mass flow rate and
equipment of reduced size.

Availability and cost:

Traditional refrigerants used in ORC are expensive. This cost could be reduced by a more
massive production of those refrigerants, or by the use of low cost hydrocarbons. The
fluid selected has to be commercially available from several suppliers at an acceptable cost.

Chemical stability:

10
Under a high pressure and temperature, organic fluids tend to decompose, resulting in mate-
rial corrosion and possible detonation and ignition. Thermal stability at elevated temperature
is thus a principle consideration in working fluid selection.

Vapour curve:

The preferred characteristic for low temperature ORC is the isentropic and positive slope
saturation vapour curve, since the purpose of the ORC focuses on the recovery of low grade
heat power, the necessary degree of superheat (to avoid wet steam at the end of turbine ex-
pansion) may not be available.

Pressures:

The maximum operating pressure required in the ORC process should be appropriately
chosen for example, high pressure processes require the use of expensive equipment
and increasing complexity but also high pressure implies high densities and hence smaller
heat exchanger and expander. Particular consideration is given to condensing pressure and
volume as they are directly related to cycle operation and maintenance and equipment size.

Vapour specific volume:

Vapour specific volume at saturation (condensing) conditions give an indication of condens-


ing equipment size. Noticing that organic fluid vapour volume varies by three orders of
magnitude between n-pentane (vapour specific volume 0.4 m3/kg at 30C)and n-dodecane
(vapour specific volume 400 m3/kg at 30C) highlights the importance of this informa-
tion in selecting the working fluid. Organic fluids with low saturation vapour volumes, like n-
pentane, require smaller condensing equipment and contribute to the choice of these working
fluids for applications where minimizing size and complexity is a priority.

Compatibility with lubricating oil:

Organic fluids must coexist with lubricating oil. The selection of a suitable oil requires care-
ful consideration of the desired physical and chemical properties, as well as the working fluid
and materials of construction to be used. Numerous investigations of the behaviour of oils in
contact with organic fluids have been conducted. In general, studies have shown that some
oils are more stable toward organic fluids than others, with increased temperature accelerat-
ing the refrigerant-oil reaction. The reaction rate is also dependent on the kinds of metal in
contact with the oil and organic fluid, the amount of air and moisture present, and the addi-
tives present in the oil.
11
Material Compatibility:

The working fluids should be non-corrosive to the more common engineering materials
used for the different components of the ORC such as pipes, heat exchangers, seals
etc.

Thermal conductivity:

A high conductivity represents a better heat transfer in heat-exchange components. The


thermal conductivity must be high in order to achieve high heat transfer coefficients
in both the employed condensers and vaporisers.

Viscosity:

The viscosity of the working fluid should be maintained low in both liquid and vapour
phases in order to achieve a high heat transfer coefficient with reduced power con-
sumption. Working fluid liquid and vapour viscosities have to be low to minimize frictional
pressure drops and maximize convective heat transfer coefficients.

Melting point:

The melting point temperature should be lower than the lowest ambient operating tem-
perature in order to ensure that the working fluid will remain in the liquid phase.

Thermodynamic performance:

The efficiency of an ORC is a well-known process performance indicator that provides a


valid assessment of the potential production of power from the process. The efficiency
and/or output power should be as high as possible for the given heat source and heat sink
temperatures.

2.4 Working fluid selection for ORC


Utilizing conventional steam Rankine cycle (RC) to recover energy from low temperature
heat sources is not the best choice technologically or the most selection economically due to
dis-satisfactory performance. With Rankine cycle engines, Badr et al [11] investigated sixty
eight potential working fluids between the operating temperature 40-120C. They concluded
that Organic fluids possess several advantages over water as the working fluid in a Rankine-
cycle engine utilising low-grade energy as the heat input, particularly for low power output
applications.

12
Numerous works in literature concentrate on fluid selection, e.g. [32-34, 36, 38]. Initially,
fluid selection recommendations were based on simple thermodynamic performance criteria.
However, in one of the earliest works about ORCs, Angelino et al. [39] already stipulated the
importance of taking technical criteria into account. More recently the working fluid selection
takes design and financial appraisal into consideration [40, 41].

Working fluids can be classified into (i) dry, (ii) isentropic or (iii) wet respectively with re-
spect to the slope of saturation curve in T-s diagram to be positive, infinite or negative [42] as
shown in Figure 2.2. Wet fluid is normally inappropriate for ORC system or the wet fluid
should be superheated to avoid liquid droplet impingent in the turbine blades during the ex-
pansion [43]. Kuo et al. [44] confirmed that there is no single physical property that can be
used as the sole indicator for quantitatively screening the working fluid and proposed a di-
mensionless figure of merit to screening the working fluid.

Hung et al [45] showed that the major physical property of screening the working fluid in-
cludes specific heat, latent heat and slope of saturation vapor curve. Tchanche et al. [46]
preferred to select working fluids with high latent heat and high specific heat. Yamamoto et
al. [47] suggested that low latent heat is better. However, Badr et al. [11], Chen et al. [48],
Hung et al. [33] suggested that high latent heat and low liquid specific heat are preferable.
Appreciable inconsistencies are encountered when screening working fluids as mentioned
above and few people contribute to analyse theoretically.

Figure 2.2 Dry, Isentropic and Wet working fluids [43]

13
Working fluid and Environmental impact

Working fluids those are generally used in ORC can be categorized according to their compo-
sition:

Chlorofluorocarbons (CFCs)
Hydro - Chlorofluorocarbons (HCFCs)
Hydrofluorocarbons (HFCs)

CFC refrigerants have been largely replaced for new equipment in developed countries by
HCFC and HFC refrigerants, in compliance with the Montreal Protocol [49]. The environ-
mental impact of a working fluid, when it escapes to the atmosphere, is not limited to
stratospheric ozone layer depletion. In fact, while all HFCs are harmless to the earths strato-
spheric ozone layer, some HFCs with large global warming potential (GWP) could contribute
significantly to climate change. HFCs were designated as greenhouse gases under the Kyoto
Protocol in 1997 [50] and they are currently targeted by efforts to reduce greenhouse gas in
most developed countries.

As a result, alternatives are sought for high GWP HFCs, such as HFC-245fa, which has a
GWP of 1030 [51]. Despite the presence of chlorine in the molecule of HCFO-1233zd-E
some studies have concluded that its ODP is an extremely small value (of 0.00034) due to its
very short atmospheric lifetime [52] as compared to saturated chlorine-containing working
fluids in current use (e.g. HCFC-123 or HCFC-22).

HFO-1336mzz(Z), also known as DR-2, is a hydrofluoro-olefin (HFO) with a GWP of 2 and


zero ODP [53]. It is non-flammable and has a favorable toxicity profile based on testing to
date. DR-2 remained chemically stable in the presence of metals up to the maximum tem-
perature tested of 250 C. DR-2 thermodynamic performance under cycle conditions
representative of potential applications was evaluated through computational modelling.
DR-2 has a relatively high critical temperature, generates relatively low vapor pres-
sures and enables high cycle energy efficiencies. It could enable more environmentally
sustainable heat pump and Rankine cycle platforms for the utilization of abundantly
available low temperature heat to generate mechanical or electrical power or meet
heating duties at higher temperatures and with higher energy efficiencies than incum-
bent working fluids [54].

14
Ryan et al. [55] studied a range of thermo-physical properties including the critical properties,
vapor pressure, liquid density, ideal gas heat capacity, liquid viscosity and surface tension of
HCFO-1233zd(E). They concluded that, it is an excellent environmentally friendly candidate
for many applications including refrigeration, foam expansion agents, and as solvents.

2.5 Bottoming ORC


The industrial waste heat which is produced in industrial processes, internal combustion en-
gines and small-scale biomass combustion plants can be used to decrease the demand of
primary energy resources.

The potential for waste heat recovery from industrial processes is enormous. For example,
about 45 % of America energy consumption is released to the atmosphere as waste heat. By
recovering the waste heat 440 million tons/year of CO2 emissions could be eliminated [56].

ORCs bottoming cycles in combined power plants have been proposed previously by several
researchers. The gas turbine engine is characterized by its relatively low capital cost com-
pared with steam power plants. It has environmental advantages and short construction lead
time. However, conventional industrial engines have lower efficiencies especially at part
load. One of the technologies adopted nowadays for improvement is the combined cycle.
Hence, it is expected that the combined cycle continues to gain acceptance throughout the
world as a reliable, flexible and efficient base load power generation plant.

Najjar [57], who analysed a combination of ORC fluids and cycle layouts that resulted in a
global combined cycle efficiency slightly below 45.2%. Invernizzi et al. [58] investigated the
possibility of enhancing the performances of micro-gas turbines through the addition of a bot-
toming organic Rankine cycle which recovers the thermal power of the exhaust gases
typically available in the range of 250300C. With reference to a micro-gas turbine with a
size of about 100 kWe, a combined configuration could increase the net electric power by
about 1/3, yielding an increase of the electrical efficiency of up to 40%.

Caresana et al. [59] studied a commercial 100 kWe micro gas turbine as a topping system; a
basic thermodynamic analysis is performed to define the principal characteristics of viable
vapour bottoming cycles. The analysis pointed to a solution adopting an Organic Rankine
Cycle (ORC) with R245fa as working fluid, due both to environmental constrains and to
technical criteria.

15
Since the temperature of the engine exhaust gas is relatively high, therefore high temperature
ORC systems should be used to gain higher thermal efficiency, as well as the working fluids
should have higher critical temperature to avoid decomposition of working fluids.

In previous studies, hydrocarbons showed good thermal performance in high-temperature


ORCs for engine waste heat recovery. Vaja et al. [60] compared benzene, R11 and R134a
used in a bottoming ORC for an internal combustion engine. The results showed that the effi-
ciency of the system was the highest with benzene. Shu et al. [61] suggested that alkane-
based working fluids were preferable for engine waste heat recovery. Siddiqi et al. [62] inves-
tigated hydrocarbons from n-pentane to n-dodecane used in ORCs of different temperature
levels.

However, one obvious disadvantage of using hydrocarbons as the working fluid in an ORC
system is their flammability and explosivity, which might limit their practical applications
due to safety concerns. Blending non-flammable components (for example, some refriger-
ants), regarded as retardants, with the hydrocarbons to suppress the flammability is a viable
solution [63].

Previous investigations provided decisive insight into energy and exergy analysis of an ORC
unit, but few literature are available regarding exergy analysis of ORC using newly devel-
oped environment friendly refrigerants when coupled with high temperature heat source such
as GT exhaust. There is utmost need for detailed energetic as well as exergetic analysis for
better understanding and improvements in ORC systems and the control of environmental
emissions.

16
3 Exergy

3.1 Exergy and its Physical Meaning


In thermodynamics, the exergy of a system is the maximum useful work possible during a
process that brings the system into equilibrium with a heat reservoir [64] . When the sur-
roundings are the reservoir, exergy is the potential of a system to cause a change as it
achieves equilibrium with its environment. Exergy is the energy that is available to be used.
After the system and surroundings reach equilibrium, the exergy is zero. Determining exergy
was also the first goal of thermodynamics. The term "exergy" was coined in 1956 by Zoran
Rant (19041972) by using the Greek ex and ergon meaning "from work" [64, 65], but the
concept was developed by J. Willard Gibbs in 1873 [66].

Definition and Aspects

Exergy is a measure of work potential or disequilibrium from the environment. The formal
definition of exergy of a system or resources can be given [67] as: Exergy is the amount of
work obtainable when some matter is brought to a state of thermodynamic equilibrium with
the common components of the natural surroundings by means of reversible processes.

In another words, Exergy is work required to raise some matter from a state of thermody-
namic equilibrium with the common components of the natural surroundings to a higher state
by means of reversible processes. So, Exergy is the minimum work needed and maximum
work obtainable in system-reference interaction.

Exergy and Reversible work

Reversible work, Wrev is the maximum amount of useful work that can be produced (or the
minimum work needs to be supplied) as the system undergoes a process between the initial
and final states. When the final state is the dead state, the reversible work equals exergy.
17
Before introducing mathematical rigour, it is important to establish some important terminol-
ogy unique to thermodynamics, or to be more specific, exergetic analysis.

Environment

The environment is defined as a very large body or medium in the state of perfect thermody-
namic equilibrium implying there is no gradients or differences involving pressure,
temperature, chemical potential, kinetic or potential energy. The environment may interact
with systems in three different ways:

Through thermal interaction as a reservoir of thermal energy at temperature T0.

Through mechanical interaction as a reservoir of unstable P0dV work.

Through chemical interaction as a reservoir of a substance of low chemical poten-


tial in stable equilibrium.

For the present work the environment has been defined at a temperature T0 = 298.15 K, and
pressure P0 = 1.01325 bar. The chemical composition of the reference environment is taken as
defined in [1]:

Table 3.1 Reference Substances [1]

Standard Partial Pressure in


Chemical Element Chemical Symbol Mole fraction in the environment Pi, 00
dry air
(bar)
Ar Ar 0.0093 0.00907
C CO2 0.0003 0.000294
D D2O (g) -- 0.00000137
H H2O (g) -- 0.0088
He He 0.000005 0.0000049
Kr Kr 0.000001 0.000000098
N N2 0.7803 0.7583
Ne Ne 0.000018 0.0000177
O O2 0.2099 0.2040
Xe Xe 0.00000009 0.000000088

18
Equilibrium

Definitions of two types of equilibria are necessary:

Restricted Equilibrium (P0, T0): At restricted equilibrium, mechanical and thermal equilib-
rium between the system and the environment is established, i.e. P = P0 and T = T0. The term
restricted implies there is a physical restriction between the system and environment prevent-
ing exchange of matter. The state of restricted equilibrium will be referred to as
environmental state.

Unrestricted Equilibrium (P00, T00, X00): In addition to mechanical and thermal equilibrium
(P = P0 = P00 and T = T0 = T00), when conditions for chemical equilibrium, i.e., equalization
of chemical potentials is established between the system and the surrounding, the unrestricted
equilibrium or the dead state is achieved.

Reference Substances

Suitable substances that present in the defined environment with known concentration, for
calculation of chemical exergy of any given element or species. Among different environ-
mental substances containing a particular chemical element the one with the lowest chemical
potential is most suitable as a reference substance.

3.2 Components of Exergy of a Thermodynamic System


Exergy associated with work transfer (Exw)

Exergy associated with heat transfer (ExQ)

Thermo-mechanical exergy (Exth)

Chemical Exergy (Exch)

3.2.1 Exergy associated with work transfer (Ex w)

Exergy associated with work transfer is equal in magnitude and expressed with the same sign
as the amount work it corresponds to. Hence

Exw = W = Wc - Wsurr ( for boundary work )


(3.1)
=W ( for other form of work )

19
Where, Wsurr = P0 (V2 - V1 ) , P0 is the atmospheric pressure and V1 and V2 are the initial and fi-

nal volumes of the system. The exergy transfer for shaft work and electrical work is equal to
the work W itself. The exergy transfer by work is zero for systems that have no work.

3.2.2 Exergy associated with heat transfer (Ex Q)

The exergy of heat transfer at the control surface is determined from the maximum obtainable
work from it using the environment as a reservoir of zero-grade thermal energy. For a heat
transfer of q and a temperature at the control surface where heat transfer is taking place T, the
maximum possible conversion from thermal efficiency to work is [5],

T
ExQ = (1 - )Q (3.2)
T0

When the source temperature is lower than that of the environment, the exergy can be viewed
as the minimum work input required in a reversible heat pump (RHP) to maintain that low
temperature, as shown in Figure 3.1.

Figure 3.1 Exergy Transfer with Heat

3.2.3 Thermo-mechanical exergy (Exth)

The thermo-mechanical exergy is equal to the maximum amount of work available when a
stream of substance is brought from its initial state to the environmental state defined by P0
and T0, by physical processes involving only thermal interaction with the environment. For
any given state, it is given by [5],

20
Exth = ( h - T0 s) - ( h0 - T0 s0 ) (3.3)

Where,

h specific enthalpy of the stream at the given state


s specific entropy of the stream at the given state
h0 specific enthalpy of the stream at the environmental state
s0 specific entropy of the stream at the environmental state

For closed-mass system, the expression is changed to [5],

Exth = (u + P0v - T0 s ) - (u0 + P0v0 - T0 s0 ) (3.4)

Where,

u specific internal energy of the stream at the given state


v specific volume of the stream at the given state
u0 specific internal energy of the stream at the environmental state
v0 specific entropy of the stream at the environmental state

Figure 3.2 Components of Exergy [68]

21
3.2.4 Chemical Exergy (Exch )

Thermodynamically, the maximum work of a chemical reaction can be put as [1],

[Wx ]max = -Dg 0 (3.5)

Where, Dg 0 is the standard value of Gibbs function of the reaction. The relevance if equation
(3.5) is not limited to chemical processes. Work in any isothermal process (e.g., diffusion) is
equal to the decrease of the Gibbs function of the stream [1].

Chemical Exergy of Reference substances

When the system contains any of the reference substances, its difference in concentration
from of the environment is the source of work potential which could be equilibrated through
diffusion. Hence, the molar chemical exergy for a reference substance is given by [1],

Exch ,rs = ( m0 - m00 ) rs (3.6)

Which under ideal gas assumptions, can be shown as [1],

P0
Exch ,rs = RT0 ln( ) (3.7)
P00

Where, the subscript rs denotes reference substance and

0 the chemical potential of the substance at environmental state


00 the chemical potential of the substance at dead state
R the gas constant
P00,rs the partial pressure of the substance in the reference environment

Chemical exergy of non-reference substances

These substances are first converted to reference substances using fictitious reversible reac-
tions involving only one mole of the reference substance while only heat transfer with the
atmosphere is allowed. Thus chemical exergy for these substances comprises of the maxi-
mum work from these reactions, minus chemical exergies of the product reference
substances. If the reference reaction takes j reference reactant species from the environment
and one mole of the non-reference substance, and produces k reference products, the maxi-
mum work of this reaction is [1],

22
-Dg 0 = ( prod ,rs nk g k0 - reac ,rs n j g 0j - g nrs
0
) (3.8)

Hence, the expression for molar chemical exergy becomes [1],

ch ,rs - Dg 0 - n ( Ex
Ex ch ,rs ) + n ( Ex ch ,rs ) (3.9)
reac j j prod k k

Where, the subscript nrs denotes non-reference substance, and


nj the number of moles of j-th reactant reference substance
nk the number of moles of k-th product reference substance

Chemical exergy of a mixture

If the substances from an ideal gas mixture, for example gaseous fuels, combustion products,
etc., the constituents of the mixture are separated by an ideal device through a reversible and
isothermal separation and compression process. The work required for these processes per
mole of the gas mixture is [1],

mix ] = RT
i [W x ln x (3.10)
rev 0 i i i

Where, xi is the mole fraction of the i-th component in the mixture. Thus exergy of the mix-
ture is equal to the sum of molar chemical exergies of the constituents plus the reversible
work that goes into separating them [1],

ch,mix = Ex
Ex x ln x
ch ,i + RT (3.11)
i 0 i i i

Investigation of equation (3.11) reveals the last summation term on the right to be negative.
Physically, it signifies the reduction in work potential of the constituents as they form a mix-
ture.

3.3 Exergy Balance of Thermodynamic Systems


With the components defined, total exergy of a system or stream is defined as the sum of the
thermo-mechanical and chemical components [69].

Extotal = Exth + Exch (3.12)

Where, for open system, Exch is calculated using equation (3.3), and for a closed system,
equation (3.4) is used instead.

23
Hence, the general exergy balance can now be expressed in extensive form using the compo-
nent definition stated in the previous subsections as,

Exi + ExQ = Exe + Exw + Ed (3.13)

Where,
Exi the initial exergy of the system or stream
Exe the final exergy of the system or stream
ExQ the exergy transfer with heat
Exw the exergy transfer with work
Ed the amount of exergy destruction (generated irreversibility)

3.3.1 Exergy Destruction (Irreversibility)

Exergy is not a conserved property. It is destroyed by the generation of entropy. The second
law provides useful relations concerning entropy generation through dissipation (e.g., fluid
friction, ohmic resistance) and spontaneous non-equilibrium processes (e.g., spontaneous
chemical reaction, free diffusion, unrestrained expansion, equalization of temperature etc.),
called in general exergetic terms, exergy destruction (irreversibilities) of a process. The
Gouy-Stodola relation provides a convenient measure of the amount of exergy destruction of
a process in an open system with multiple inlets and outlets, and a number of thermal interac-
tions at different temperatures as [1],

Q
Ed = T0 me se - mi si - r (3.14)
out in r Tr

For a closed mass system, which can be modified as [1],


Q
Ed = T0 DS - r (3.15)
r Tr

Where, Tr is the temperature at the system boundary with a heat interaction Qr.

3.3.2 Second Law Efficiency

A performance parameter based on the exergy concept is known as the Second Law Effi-
ciency (II) or as Second Law Effectiveness ( ).A first-law efficiency gages how well the
energy is used when compared against an ideal process, whereas an effectiveness indicates

24
how well exergy is utilized. Exergetic efficiency or, effectiveness is defined in general sense
as the ratio of the sum of all output exergy, Exoutput over the sum of all input exergy, Exinput
[5],

Exoutput Exergy destruction


= = 1- (3.16)
Exinput Exergy in

As it considers the quality of energy using the exergy concept, thus this efficiency definition
is superior to the conventional one. And it allows more sensible direction in energy auditing,
decision making and thermodynamic design of power systems.

25
4 Thermodynamic Modelling

4.1 Thermodynamic Modelling of ORC system


The thermodynamic system analysis consists of three statements concerning three system
properties: mas, energy and entropy. These encapsulated the mass conservation law, the en-
ergy conservation law (first law of thermodynamics), and the exergy (second law of
thermodynamics).

Exergy is not conserved as energy, which is destructed in the system due to internal and ex-
ternal irreversibilities. For a real process, the exergy input always exceeds the exergy outputs;
this unbalance is due to irreversibilities, a process known as exergy destruction. Thus, ther-
modynamic inefficiencies and the processes that cause them are identified. Illustration of the
above three indifferent mathematical forms can be shown as follows:

Mass Balance

The mass rate balance for control volumes with several inlets and exits which is commonly
employed in engineering is [5],

dmcv
= i m i - e m e (4.1)
dt

Where, the subscripts


cv the control volume
i inlet
e outlet
m mass flow rate of the fluid stream
dmcv
For steady state, = 0. So equation (4.1) becomes [5],
dt
i m i - e m e = 0
(4.2)
i m i = e m e

Energy Balance

An accounting balance for the energy of the control volume is [5],


dEcv V2 V2
= Q cv - Wcv + i m i (hi + i + gzi ) - e m e (he + e + gze ) (4.3)
dt 2 2
26
Where,
dEcv
the time rate of change of energy of control volume
dt
Q cv time rate of heat output in the control volume

W cv time rate of work output in the control volume

h enthalpy
V bulk velocity of the working fluid
z altitude of stream above the sea level
g specific gravitational force

Exergy balance

The control volume exergy rate balance considering several inlets and exits in control volume
is [5],

dExcv T dV
= i (1 - 0 )Q j - (Wcv - p0 cv ) + i m i e f i - e m e e f e - E d (4.4)
dt Tj dt
rate of change Exergy transfer
Net exergy outflux
of exergy of Exergy transfer with mass flow
with work
control volume with heat

Where

Q j Time rate of heat transfer at the location on the boundary where the instanta-
neous temperature is T j

W cv Time rate of energy transfer by work other than flow work

E d Net exergy destruction rate within CV

Flow exergy (ef) of a fluid in steady flow is defined as the maximum work output that can be
obtained as the fluid is changed reversibly from the given state to a dead state in a process
where any heat transfer occurs solely with the environment The specific flow exergy is given
by [5],

V2
e f = (h - h0 ) - T0 ( s - s0 ) + + gz (4.5)
2

Neglecting kinetic and potential energy [5],

e f = ( h - h0 ) - T0 ( s - s0 ) (4.6)

27
4.2 Mathematical Model

4.2.1 Energy and Exergy analysis of the cycle components

In this section, individual components energy and exergy balance of the cycles are intro-
duced under the assumed conditions of the present study.

4.2.1.1 Energy Analysis


To analyse the possible realistic performance, a detailed energy analysis of the topping gas
turbine cycle and bottoming ORC has been carried out by ignoring the kinetic and potential
energy balance. For steady state flow the energy balance for the thermal system can be writ-
ten modifying the equation (4.3) as below [5]:

V2
V2
0 = Q cv - Wcv + i m i (hi + i + gzi ) - e m e ( he + e + gze ) (4.7)
2 2

Neglecting kinetic and potential energy [5],

0 = Q cv - Wcv + i m i hi - e m e he (4.8)

For one inlet and one exit system [5],

0 = Q cv - Wcv + m i hi - m e he (4.9)

0 = Q cv - Wcv + m (hi - he ) (4.10)

For single component analysis, equation (4.10) will be used in this study.

The energy or first law efficiency h I of a system and/or system component is defined as the

ratio of energy output to the energy input to system/component [5], i.e.,

Desired Output Energy


hI = (4.11)
Input Energy Supplied

4.2.1.2 Exergy Analysis


Exergy Analysis is a method that uses the conservation of mass and conservation of energy
principles together with the second law of thermodynamics for the analysis, design and im-
provement of energy systems. The exergy method is a useful tool for furthering the goal of
more efficient energy-resource use. Many engineers and scientists suggest that the thermody-
namic performance of a process is best evaluated by performing an exergy analysis in

28
addition to or in place of conventional energy analysis because of exergy analysis appears to
provide more insights and to be more useful in furthering efficiency improvement efforts than
energy analysis [70].

The exergy (second law) analysis has become one of the significant methods to evaluate the
performance of any thermal system application because the exergy analysis deals with the
quality of energy. Exergy can be defined as availability, the highest available work, which is
an evaluation of the maximum useful work that can be obtained when a system is brought to
a state of equilibrium with the environment in a reversible process. For a real process, the ex-
ergy input always exceeds exergy outputs, this unbalance is due to irreversibilities, which is
known as exergy destruction.

A general form of the exergy equation for an open system control volume already stated in
equation (4.4). For steady state flow the exergy balance for a thermal system is given in equa-
tion (4.12), where time rate variations are neglected [5].

T0 dV
0 = i (1 - )Q j - (Wcv - p0 cv ) + i m i e f i - e m ee f e - E d (4.12)
Tj dt

Rearranging equation (4.12) gives the exergy destruction of a steady state open system for a
control volume [5].

T dV
E d = i (1 - 0 )Q j - (Wcv - p0 cv ) + i m i e f i - e m ee f e (4.13)
Tj dt

If the system is adiabatic, then Q j = 0 . So, for a steady state adiabatic system equation (4.12)

becomes [5],

0 = -Wcv + i m i e f i - e m ee f e - E d (4.14)

So, for 1-inlet, 1-exit system [5],

0 = -Wcv + m (e f i - e f e ) - E d
(4.15)
E d = m (e f i - e f e ) - Wcv

Total exergy of a system consists of four different components:

Extotal = Ex ph + Exkn + Ex pt + Exch (4.16)

Neglecting the potential and kinetic exergy, equation can be written as

29
Extotal = Ex ph + Exch (4.17)

4.2.2 Thermodynamic Relationships in the ORC

The schematic representation of the ORC is illustrated in Figure 4.1. In the system, working
fluid is pumped, firstly. Namely, low pressure fluid is compressed to high pressure fluid by a
pump (state 5 to 6). Then high pressure fluid enters and passes through the evaporator. In the
evaporator, high pressure fluid has become heated and pressurized vapor using the heat ca-
pacity of inlet water (state 6 to 3). After that the heated and pressurized vapor enters in
turbine. And it leaves from turbine as low pressure vapor and generates electricity (state 3 to
4). Lastly, the low pressure vapor goes through the condenser, and the working fluid leaves
from condenser as saturated liquid (state 4 to 5) and the cycle continues [71].

Figure 4.1 Schematic Illustration of the ORC

4.2.2.1 Energy balance equations for ORC


Heat load absorbed by the ORC [5],

Q HS = m hs .C p , HS .(T1 - T2 )
(4.18)
or , Q HS = m hs (h1 - h2 )

30
Evaporator: (Process 6 to 3)

Figure 4.2 Schematic of ORC evaporator

Heat transfer rate in the evaporator, by neglecting any heat loss to the surroundings, can be
expressed as [5],

Q eva = m hs (h1 - h2 ) = m r ( h3 - h6 ) (4.19)

Pump: (Process 5 to 6)

Figure 4.3 Schematic of ORC pump

If the pump is adiabatic, then Q = 0 . So pump work can be written as [5],

m (h - h )
W p = r 5 6 s (4.20)
h pump

Turbine: (Process 3 to 4)

Figure 4.4 Schematic of ORC turbine

31
For an adiabatic turbine, Q = 0 . So, turbine work [5],

Wt = m r (h3 - h4 s ).hturb (4.21)

Condenser: (Process 4 to 5 and 7 to 8)

Figure 4.5 Schematic of ORC condenser

Heat transfer rate in the condenser, by neglecting any heat loss to the surroundings, can be
expressed as [5],

Q con = m r ( h4 - h5 ) = m cw (h8 - h7 ) (4.22)

4.2.2.2 Exergy balance equations for ORC


Exergy balance equations are derived by modifying equation (4.14).

For any particular point, e.g. point i, the specific flow exergy is found by equation (4.6),

e fi = (hi - h0 ) - T0 ( si - s0 ) (4.23)

Heat source exergy in:

Heat Exergy in from heat source [5]:


= m (e - e ) = m [(h - h ) - T ( s - s )]
Ex (4.24)
in hs f1 f2 hs 1 2 0 1 2

Evaporator:

Exergy balance equation for evaporator is obtained from equation (4.14) [5],
m hs (e f 1 - e f 2 ) + m r (e f 6 - e f 3 ) - E d ,eva = 0
(4.25)
E d ,eva = m hs (e f 1 - e f 2 ) + m r (e f 6 - e f 3 )

32
The second law efficiency [5],
m r (e f 3 - e f 6 )
h II ,eva = (4.26)
m hs (e f 1 - e f 2 )

Pump:

Exergy balance equation for pump is obtained from equation 5.5 ,

0 = -W p + m (e f 5 - e f 6 ) - E d , pump
(4.27)
E d , pump = m (e f 5 - e f 6 ) - W p

The second law efficiency [5],


m r (e f 5 - e f 6 )
h II , pump = (4.28)
W p

Turbine:

Exergy balance equation for turbine is obtained from equation 5.5 ,


0 = -Wt + m (e f 3 - e f 4 ) - E d ,turb
(4.29)
E d ,turb = m (e f 3 - e f 4 ) - Wt

The second law efficiency [5],


m r (e f 3 - e f 4 )
h II ,turb = (4.30)
W t

Condenser:

Exergy balance equation for evaporator is obtained from equation (4.14),

m r (e f 4 - e f 5 ) + m cw (e f 7 - e f 8 ) - E d ,cond = 0
(4.31)
E d ,cond = m r (e f 4 - e f 5 ) + m cw (e f 7 - e f 8 )

The second law efficiency [5],


m cw (e f 8 - e f 7 )
h II ,cond = (4.32)
m r (e f 4 - e f 5 )

4.2.2.3 Overall System Analysis


Net work done [5]:
Wnet = Wt - W p (4.33)

33
Heat energy in [5]:
qin = m hs (h1 - h2 ) (4.34)

Thermal Efficiency/1st law Efficiency [5]:


Wnet
h I ,sys = (4.35)
qin

Second Law Efficiency [5]:


W
h II ,sys = net (4.36)
Ex ,in

Thermal efficiency and second law efficiency are key performance criteria of ORC, because
they quantify how much of the input energy/exergy is being transformed into desired output.
Based on the thermodynamic model developed in this chapter, proper simulation is per-
formed. The simulation results are the primary topics of the next chapter.

34
5 Simulation and Results

Methodology which was discussed above is used to analyse the performance of the basic
ORC. In the thermodynamic analysis, each component in the system are modelled. This chap-
ter starts with the validation of the model with the available experimental data. Next, suitable
organic fluids are selected based on thermal condition of the heat source and environmental
impact. Selected fluids are tested one by one and variations of key performance parameters
such as net power output, power consumption, mass flow rate, energy and exergy efficien-
cies, exergy destruction of the system as well as components are analysed. In the exergy
analysis, component wise exergy destruction/losses are evaluated to determine distribution of
the exergy destruction in the systems. This chapter presents the detailed explanation of the
findings of the variables described above in a chronological order.

5.1 Validation of the Model developed


A Bottoming Organic Rankine Cycle (ORC) is modelled using governing thermodynamic
relations. To validate the model, the results have been compared to the available experimental
data.

The thermodynamic properties has been acquired using COOLPROP [72] and the simula-
tion of the ORC system has been performed by a computer program written in the
PYTHON environment. To handle combustion, chemical equilibrium solver of
CANTERA [73] module has been used. Canteras chemical equilibrium solver is handy to
set a gas mixture to a state of chemical equilibrium holding temperature and pressure fixed.

5.1.1 Validation of Topping Cycle (GT Cycle)

A gas turbine model is developed using Python environment and the model is validated with
the results of P.S. Pak [74]. Both energy and exergy balance data are tabulated in Table 5.1
and Table 5.2 respectively, followed by a schematic representation of energy balance in Fig-
ure 5.1. Gas Turbine inlet temperature is considered 1273 K. Here, the results of the model of
the present study show good agreement with those of the experimental results of Literature
[74].

35
Table 5.1 Energy Balance (MW) Comparison with System A of [74]

Item Literature [74] Present Study


Lower Heat value of fuel 36.97 36.94
Generated Power Output 10.0 10.28
(Turbine Output) 32.23 32.24
(Power for compressing air) 21.7 21.43
(Power for com fuel gas) 0.507 0.52
Heat Output 16.93 16.9
Total of effective used heat 26.91 27.2
Flow loss at air compressor 0.22 0.22
Heat loss at waste heat boiler 0.887 0.82
Heat of ambient exhaust gas 6.65 7.10
Generator Loss 0.22 0.28

Figure 5.1 Energy balance Comparison with System A of Literature [74]

Table 5.2 Exergy Balance (MW) Comparison with System A of [74]

Item Literature [74] Present Study


Generated Power Output 10.017 10.28
(Turbine Output) 32.26 32.24
(Power for compressing air) 21.72 21.43
(Power for com fuel gas) 0.51 0.52
Exergy of supplied Heat 5.22 5.73

36
Total of effective used exergy 15.21 16.01
Irrev. loss at filter silencer 0.073 0.087
Flow loss at air compressor 0.144 0.20
Irrev. loss at air compressor 1.43 1.34
Irrev. loss at fuel compressor 0.036 0.041
Irrev. loss at combustor 11.27 11.62
Irrev. loss at turbine 2.50 2.46
Exergy of ambient exhaust gas 1.105 1.082
Generator loss 0.221 0.285

5.1.2 Validation of Bottoming ORC model

Optimum evaporation temperatures for any specific pinch point temperature difference and
any heat source outlet temperature are obtained by simulation. The results show good agree-
ment (Table 5.3) with those found by the model of J.Song et al. [75].

Table 5.3 Comparison between the results of [75] and Present Study

Optimum Evaporation
Ref. flow rate, kg/s
Heat Working Temperature (K)
Source fluid Literature Present Literature Present
[75] Study [75] Study
System C R236ea 381.5 382.75 43.42 43.54
(147 C) R245fa 375.0 375.5 37.04 36.77
System D R141b 352.2 352.25 45.21 45.18
(104 C) n-Heptane 352.9 353 25.97 25.962
System E R141b 350.9 350.75 25.63 25.64
(98 C) 2- Butene 351.4 351.5 14.90 14.892

After finding out the optimum evaporation temperatures, several analyses are carried out by
varying the outlet temperature of heat sources. The results uphold the findings of O. Kaska
[26] ( Figures 5.2 - 5.5 ).

37
Figure 5.2 Comparison with refrigerant flow rates of O. Kaska [26]

Figure 5.3 Comparison with heat available of O. Kaska [26]

38
Figure 5.4 Comparison with turbine power of O. Kaska [26]

Figure 5.5 Comparison with pump power of O. Kaska [26]

39
From the above comparison results, it is evident that the results obtained from the ORC
model of this present study show satisfactory resemblance with those of the previous models,
which substantiates the validity of the model developed.

5.2 Modelling Assumptions


The following assumptions are made in the model to simplify the analysis of the system:

All processes of the cycle are in steady state.


Fuel for the gas turbine is supposed to be pure methane and its temperature is
equal to the ambient temperature.
The mass flow rate of the GT exhaust gas entering the ORC is 50 kg/s.
All components are assumed to operate under adiabatic conditions.
The isentropic efficiency of the compressor and turbine of the ORC are as-
sumed to be constant and equal to 75%.
The pinch point temperature difference between the flue gas and organic fluid
is assumed as 10 C.
The condensing temperature of the working fluid is assumed as 40 C.
Working fluid exits the condenser as saturated liquid (x = 0).
There is 5% pressure difference between the flue gas entering and exiting the
evaporator of the ORC.
The working fluid enters the turbine of the ORC as superheated vapor.
Temperature rise of the condenser cooling water is 10 C.
Dead state temperature and pressure are 298.75 K and 101.325 kPa respec-
tively.
Pure counter current flow in heat exchangers.
Changes in the kinetic and potential energy of the internal and external fluid
streams are negligible.
Heat losses in the heat exchangers are negligible (2%). Hence, the rate of heat
transfer to the working fluid is almost equal to that extracted from the heat
source.

40
5.3 Screening of Potential working fluids
The selection of working fluid has a great effect on the operating condition, system effi-
ciency, economic viability and environmental impact. Regarding the environmental issues,
several low GWP organic fluids capable of sustaining at higher temperature are selected for
the analysis. From an environmental perspective, switching from conventional refrigerants to
low GWP refrigerants is the most important consideration.

Properties of screened working fluids have been tabulated in Table 5.4 according to their
GWP values. R245fa is a popular refrigerant for ORC with higher GWP values and listed to
compare the performance with the other low GWP fluids.

R1233zd(E) [trans-1-Chloro-3,3,3-trifluoropropene] and R1234ze(Z) [cis-1,3,3,3-


Tetrafluoroprop-1-ene] are being investigated as a working fluid possessing a low global
warming potential (GWP) for high-temperature heat pumping applications, Organic Rankine
Cycles, and air-conditioning and refrigeration applications [76]. According to the brief risk
assessment conducted by Koyama et al. [77], the flammability of R1234ze(Z) seems to be
very mild, and the acute inhalation toxicity is very likely negative. Fukuda et al. [78] numeri-
cally quantified the irreversible loss and suggested that R1234ze(Z) is likely beneficial to
high-temperature heat pumps. However, the heat transfer data of these new refrigerants for
such high-temperature conditions is not yet available in the open literature. In addition to the-
se two very low GWP fluids, several moderate GWP fluids (R141b, R124, R21) have been
investigated in this study. All the selected working fluids are one-component and mixtures
are not included in this study because mixing rules are rather complicated.

Table 5.4 Properties of Screened Organic Fluids

Critical Critical
Molar mass
Refrigerants Temperature Pressure GWP
(kg/mol)
(K) (bar)
R245fa 427.01 36.51 0.134 1030
R141b 477.50 42.12 0.116 725
R124 395.42 36.24 0.136 609
R21 451.48 51.81 0.102 151
R1233zd(E) 438.75 35.70 0.130 0
R1234ze(Z) 423.27 35.33 0.114 0
Water 647.12 220.64 .018 --

41
5.4 Drawback of water over organic fluids
Organic fluids possess several advantages over water as the working fluid in a
Rankine Cycle engine utilising low-grade energy as the heat input, particularly for low
power output applications [11]. Thermodynamic performance in utilization of low grade
heat source is regulated by the slope of T-s curve of the saturated vapor. Water which exhibits
negative slope (dT/ds < 0) of the saturated vapor curve (Figure 5.7), is not suitable for low
grade heat sources, because after expansion through turbine, the quality of steam is very low
(Figure 5.6). Presence of water droplets is detrimental to turbine blades causing serious tur-
bine blade damage.

Organic fluids have inherent advantage in this case because of the slope of saturated vapor
curve (Figure 5.8) is either positive (dry fluids) or infinite (isentropic fluids). There is no pos-
sibility of presence of droplets in the working fluid after the expansion through the turbine.
Superheating is not necessary even at very low temperatures, thus, preventing the turbine
blade erosion. So, organic fluids are suitable in this case instead of water. For the advantages
of dry fluids over wet fluids, wet fluids have been discarded in this analysis.

Figure 5.6 Comparison of wet fluid and dry fluid [79]

42
Figure 5.7 T-s diagram of water

Figure 5.8 T-s diagram of several organic fluids

43
5.5 Parametric Effect on the performance of ORC
Since water is not suitable as an ORC fluid, in this section, performance of ORC is ana-
lysed using selected refrigerants as working fluid. Effect of evaporator pressure for a fixed
turbine inlet temperature on various performance parameters such as turbine work output,
power consumption of pump, mass flow rate, thermal efficiency, exergy efficiency, irreversi-
bilities of individual components are discussed in this section. Selected working fluids are
tested one by one and finally comparative analyses are made.

However, initially, for a particular working fluid R245fa, variations of all the above-
mentioned parameters with changing evaporator pressure are discussed. R245fa is chosen
here because the validation of the ORC model was carried out using R245fa as the working
fluid. It is commonly used in many operating ORC systems and seems to be a promising
working fluid. R245fa possesses its own advantage in the operation of the low temperature
ORC system.

Table 5.5 shows various representative energy and exergy performance data for ORC pa-
rameters at evaporator pressure 10 bar obtained from the simulation of the model developed
for refrigerant R245fa.

Table 5.5 Representative Energy and Exergy performance data for R245fa at 10 bar

Overall Heat
Total Exergy 1 st Law Exergetic
Efficiency (%) Transfer
Energy Component Destruction Efficiency Efficiency
1 st 2 nd or Power
In (kW) (kW) (%) (%)
Law Law (kW)
Evaporator 20450 1681 98.0 80.1
Turbine 1383 220 75.0 86.5
Condenser 18746 4402 98.0 12.3
20929 7 17 Pump 26 7 75.0 76.2
Exergy re-
jection at - 613 - -
condenser

In the beginning of the analysis, complete energy and exergy balance of the cycle will be
made to substantiate the validity of the model for a certain evaporator pressure 10 bar, fol-
lowed by the analysis of various performance parameters.

44
5.5.1 Energy and Exergy Balance of ORC

Energy balance for the complete system is represented in Figure 5.9 and Figure 5.10 at a par-
ticular evaporator pressure 10 bar. From the energy balance diagram, it is clear that a major
portion of the net energy input is dissipated as rejected heat to cooling water. Power con-
sumption of the pump is relatively lower as expected since refrigerant enters the pump as
saturated liquid. The total energy input is distributed to the net power output, condenser rejec-
tion and heat loss at the evaporator and condenser. In accordance with the conservation of
energy principle, the total energy enters the control volume equals the total energy that exits
in various forms. However, the total rate of exergy enters the control volume exceeds the total
rate of exergy exits. The difference between these exergy values is the rate at which exergy is
destroyed by irreversibilities, in accord with the second law.

Figure 5.11 shows the exergy balance for the complete system as well as the major contribu-
tors of the exergy destruction. Although a major portion of energy is released as heat in the
condenser, from the figure it is clear that its exergy value is very much small, only 7% of the
total exergy input. As much as 17% of the total exergy is received as work.

Figure 5.9 Component-wise energy consumption for R245fa at 10 bar evaporator pressure

45
Figure 5.10 Energy Balance for the system for R245fa at 10 bar evaporator pressure

Figure 5.11 Exergy Balance for the system for R245fa at 10 bar evaporator pressure

The rest 76% is lost due to irreversibility of the system, also known as exergy destruction. Of
all components, condenser is the primary source of exergy destruction contributing around
70% of total exergy destruction. Exergy destruction in the evaporator, turbine and pump are

46
26%, 3.9% and 0.1% respectively. Although heat exchangers (e.g., evaporators and condens-
ers in this case) appear from an energy perspective to operate without any loss when stray
heat transfer is ignored, they are primary sites of thermodynamic inefficiency quantified by
exergy destruction. The high temperature difference between the heat exchanging streams is
an indicator of heat transfer irreversibility; condenser itself is responsible for 70% of the total
exergy destruction in this case.

5.5.2 Exergetic efficiency of components at 10 bar

The exergetic efficiency provides a true measure of the performance of an energy system
from the thermodynamic viewpoint. Figure 5.12 shows exergetic efficiency of the compo-
nents as well as the overall exergetic efficiency of the system for evaporator pressure 10 bar.
Any device with higher exergy destruction is likely to have lower exergetic efficiency or vice
versa as evident in the Figure 5.12.

As stated in previous section, condenser results in higher exergy destruction, so it ends up


having the lowest exergetic efficiency among other components. Evaporator, turbine and
pump exhibit moderate exergy efficiency from the thermodynamic viewpoint. Low exergy
efficiency of the condenser also responsible for the lower overall efficiency of the system as
indicated in the figure.

Figure 5.12 Exergy Efficiency of components for R245fa at 10 bar evaporator pressure

47
5.6 ORC performance analyses at different evaporator pressures for R245fa
Now that ORC performance at a particular evaporator pressure of 10 bar has been described,
its worth analysing the ORC performance by varying evaporator pressure keeping other con-
ditions same as the previous section. For a particular working fluid R245fa, several ORC
parameters obtained by simulation have been tabulated in Table 5.6. For better understanding
of the variations, changes of each parameter have been graphically described in the following
sections.

Table 5.6 ORC parameters at different evaporator pressures for R245fa

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kg/s) (%) (%)
(kW) (kW)
10 33.83 26 1383 449 6.6 16.7
15 33.92 44 1760 439 8.4 21.3
20 34.01 61 2018 433 9.7 24.4
25 34.10 79 2211 428 10.6 26.8
30 34.19 96 2363 424 11.3 28.6
35 34.28 114 2488 421 11.9 30.1
40 34.37 132 2593 418 12.5 31.4
45 34.47 150 2683 416 12.9 32.5
50 34.56 168 2761 413 13.3 33.4
55 34.65 186 2829 411 13.6 34.3
60 34.74 204 2889 410 13.9 35.0
65 34.84 222 2942 408 14.2 35.7

5.6.1 Net Power Output and Power Consumption for R245fa

Figure 5.13 shows gross turbine work, power required by pump and net work output of the
ORC with variation of evaporator pressure using R245fa as working fluid. It can be observed
apparently that both turbine and pump work increase gradually with increasing pressure. As a
result, net power output also increases gradually as indicated by the dotted line. Condensing
temperature has been selected as 40 C, for which the saturation pressure for R245fa is 2.51
bar. So a lower evaporator pressure, e.g. 10 bar, the expansion of the working fluid against
the turbine is small, which results in a lower power output than a higher evaporator pressure
as evident in Figure 5.13. Although power consumption by pump gradually increases with
48
pressure, the power requirement of the pump is significantly smaller than the turbine output
for any evaporator pressure.

Figure 5.13 Work Output and Consumption with evaporator pressure for R245fa

5.6.2 Mass flow rate for R245fa

Variation of mass flow rates of working fluid and condenser cooling water is described in
Figure 5.14. As evident in the figure, there is no appreciable variation of refrigerant flow rate;
approximately 3% increase in the flow rate between the evaporator pressure of 10 bar and 65
bar. In contrast, flow rate of condenser cooling water decreases gradually as evaporator pres-
sure increases.

The reason behind these changes may be attributed to the variations of temperature of the
working fluid at the exit of pump and turbine. As shown in Figure 5.15, there is no significant
variation of pump outlet temperature. However, working fluid temperature at the turbine exit
decreases appreciably with increasing pressure due to greater expansion through the turbine.
That means the amount of heat rejected at the condenser will also reduce. But since the tem-
perature increase of the condenser cooling water is fixed at 10C; according to the equation
5.2, for a constant specific heat, the mass flow rate of cooling water is supposed to decrease
with increasing pressure, which conforms to the data plotted in Figure 5.14(b).

Q rejected = m cwC p DT (5.1)


49
Figure 5.14 Refrigerant and Coolant flow rate with evaporator pressure for R245fa

Figure 5.15 Exit temperature variation with evaporator pressure for R245fa

5.6.3 Overall Efficiency of ORC for R245fa

Thermal and exergy efficiency variation of the overall cycle with evaporator pressure has
been shown in Figure 5.16. Both thermal and exergy efficiency increases gradually with
evaporator pressure. Since evaporator pressure is increased keeping the condensing pressure
fixed, so it is expected to obtain a higher efficiency at elevated pressures. Magnitude of the
efficiency values may be lower because of high condenser rejection, but it should be kept in
mind that the ORC is operating from the exhaust gas of the topping GT cycle, which would
otherwise be rejected as stray heat to the atmosphere.

50
Figure 5.16 Variation of Overall Efficiencies of ORC for R245fa

5.6.4 Irreversibility in components for R245fa

One of the important uses of the second law of thermodynamics in engineering is to deter-
mine the best theoretical performance of systems. By comparing actual performance with the
best theoretical performance, insights often can be gained into the potential for improvement.
Component-wise irreversibility analysis will enable to pinpoint the location of losses from
thermodynamics viewpoint and one way to do this is to determine the exergy destruction of
components due to irreversibility. Component-wise exergy destruction and exergy rejection at
the condenser have been tabulated in Table 5.7 for working fluid R245fa. For better visualisa-
tion of the tabulated data, a figure is plotted (Figure 5.17) to analyse the distribution of
exergy destruction of components as well as total exergy destruction of the system with in-
creasing evaporator pressure. Although exergy destruction behaviour of the components
varies differently, it is evident from the Figure 5.17 that total exergy destruction of the overall
system decreases with the evaporator pressure; total exergy destruction is approximately 20%
less at 65 bar than that of 10 bar. Quantity of exergy rejected at condenser (which is much
smaller than the exergy destructed in the same device) decreases gradually with the evapora-
tor pressure, as evident in Table 5.7.

51
Table 5.7 Exergy Destruction of components for R245fa

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
` Turbine Condenser Pump

` 1641 220 4403 6 614


15 1415 286 4193 10 601
20 1261 334 4044 14 592
25 1147 372 3929 19 586
30 1058 403 3834 23 580
35 986 430 3753 27 576
40 926 453 3683 31 572
45 877 474 3620 35 568
50 834 493 3563 40 565
55 798 510 3511 44 563
60 767 526 3463 48 560
65 740 541 3418 52 558

Figure 5.17 Distribution of exergy destruction with evaporator pressure for R245fa

52
Exergy destruction in the pump is very negligible compared to other components and varia-
tion of exergy destruction values do not change that much with changing evaporator pressure
as shown in Figure 5.17.

If thermodynamic average temperatures for the hot and cold streams exchanging heat in a
counter-flow heat exchanger are Th and Tc respectively, then at steady state the closed system
exergy rate equation can be written as [80],

T0 T
0 = (1 - )Q - (1 - 0 )Q - E D (5.2)
Th Tc

Reducing the expression gives [80],

T -T
E D = T0Q h c (5.3)
ThTc

Since the heat transfer rate is proportional to the temperature difference between the streams
[80]:
Q (Th - Tc ) (5.4)

So, we have qualitatively [80],


(T - T ) 2
E D T0 h c (5.5)
ThTc

Therefore, the rate of exergy destruction associated with heat transfer varies quadratically
with the stream to stream temperature difference and inversely with the product of the tem-
perature levels. This interprets mathematically why the exergy destruction at the condenser is
much higher in the bottoming ORC analysed in this study.

5.6.5 Exergy Efficiency variation of ORC components for R245fa

An important use of exergy efficiencies is to assess the thermodynamic performance of com-


ponent, plant, or industry relative to the performance of similar components, plants, or
industries. Variation of exergy efficiency of ORC components has been represented graphi-
cally in Figure 5.18.

53
Figure 5.18 Exergy Efficiency of the ORC components for refrigerant R245fa

Because the turbine inlet temperature is very high due to high exhaust temperature of the GT,
therefor after expanding in the turbine, the working fluid leaves at a higher temperature as
shown in Figure 5.19. But as the evaporator pressure increases, the scenario changes gradu-
ally. With the increase of evaporator pressure, pressure drop across the turbine increases,
resulting in the gradual decrease in the turbine-exit temperature as well (T 4 < T4 < T4 ). This
might account for the higher turbine work output with increasing evaporator pressure as evi-
dent in Figure 5.13.
Higher turbine exit temperature of the refrigerant is also the reason behind the greater con-
denser-rejection (Table 5.7). The hot refrigerant must reject heat in order to start a new cycle.
The condensing temperature is assumed as 40 C. Therefore, the heat rejection is colossal in
this case. Higher temperature difference between the turbine-exit and condensing temperature
is also responsible for larger irreversibility (exergy destruction) in the condenser among all
other components of the ORC (Figure 5.17).
54
Figure 5.19 Variation of performance parameters with evaporator pressure

5.7 Performance of ORC for various refrigerants


In this section, performance of ORC will be discussed by varying the refrigerants listed in
Table 5.4. Condensing temperature is assumed as 40 C. However, saturation pressure corre-
sponding to this temperature varies with the refrigerants. Since the turbine inlet temperature
of ORC is fixed, this saturation pressure variation of these refrigerants is a major contributor
to the performance variations. Table 5.8 shows variation of saturation pressure of the selected
refrigerants at condensing temperature 40 C.

Table 5.8 Saturation pressure variation of refrigerants at condensing temperature

Saturation Pressure at Condensing


Refrigerants temperature (40 C)
(bar)
R245fa 2.51
R141b 1.33
R124 5.93
R21 2.96
R1233zd(E) 2.15
R1234ze(Z) 2.90

55
5.7.1 Net Power Output

Figure 5.20 shows the effect of evaporator pressure on net work output using the selected re-
frigerants as working fluid. For all the refrigerants, net work output increases with the
evaporator pressure. However, among the working fluids, R124 results in the lowest net work
output while R141b is the greatest. This might be due to the saturation pressure of the work-
ing fluids at condensing temperature as shown in Table 5.8. R141b has the least value of the
saturation pressure, so greater expansion through the turbine. Since pump power consumption
of ORC for various refrigerants does not vary much and magnitude of these values is in the
kilowatt range (Figure 5.21), therefore net power output does not get affected that much for
the pump consumption. As a result, R141b shows the highest net work output. R124 has the
highest saturation pressure of 5.93 bar, so for the same evaporator pressure, the ORC output
should be lower. Net power output using R1233zd(E) and R1234ze(Z) is almost same, ap-
proximately 7% difference on average. Work output using R245fa is lower than all the
working fluid used, except R124.

Figure 5.20 Net work output variation for selected refrigerants

56
Figure 5.21 Pump consumption variation for selected refrigerants

5.7.2 Thermal Efficiency

For a power cycle, thermal efficiency indicates the extent to which the energy added by heat
is converted to net work output. Since temperature of the GT exhaust gas is assumed fixed,
there is not much variation in the energy added by heat. Thermal efficiency is the ratio of net
work output to energy added by heat. So working fluid resulting in higher net work output
should also result in higher thermal efficiency of ORC, which is shown in Figure 5.22. As
expected, R141b results in the highest thermal efficiency. Sequence of the refrigerants ac-
cording to the thermal efficiency from higher to lower is: R141b > R21 > R1233zd(E) >
R1234ze(Z) > R245fa > R124.

The difference in the thermal efficiencies can be explained by pressure ratio variation of the
refrigerants. For instance, at 60 bar evaporator pressure, the pressure ratio is 60/1.33 i.e.
45.12 for R141b while this value is 60/5.93 i.e. 10.11 for R124. Thats why thermal effi-
ciency values are much higher for R141b than R124. Thermal efficiency values for other
refrigerants can be described similarly using their respective pressure ratio values.

57
Figure 5.22 Thermal efficiency variation for selected refrigerants

5.7.3 Heat rejection and flow rate of coolant at Condenser

As illustrated in Figure 5.23, heat rejection to condenser cooling water decreases with the
evaporator pressure. However, there is significant variation in the rejected heat for different
working fluids. R124 results in the highest heat rejection while R21 & R141b result in the
lowest values. The reason behind these variations may be attributed to the temperature of the
working fluid at the turbine exit, which is shown in Figure 5.24.
For all refrigerants, turbine exit temperature is abated with the evaporator pressure, the reason
behind this abatement is described in Figure 5.19. Working fluids exiting the turbine at higher
average temperature tend to release more heat at the condenser whose temperature is fixed at
40 C. For instance, using refrigerant R124, the working fluid temperature at the turbine exit
varies between 330 to 385 C, which is significantly higher than that of other refrigerants.
Consequently, it releases the greatest amount of heat.
Due to constant temperature rise of 10 C of the condenser cooling water, the mass flow rate
of the cooling water varies accordingly with the condenser heat rejection. This variation is
demonstrated in Figure 5.25. Cooling water flow rate follows the same sequence as that of
the condenser heat rejection. The sequence of the working fluids according to heat rejection
or coolant flow rate is: R124 > R245fa > R1234ze(Z) > R1233zd(E) > R21 > R141b.

58
Figure 5.23 Condenser rejection variation for selected refrigerants

Figure 5.24 Expander outlet temperature variation for selected refrigerants

59
Figure 5.25 Coolant flow rate variation for selected refrigerants

5.8 Cycle Irreversibility Analysis

5.8.1 Second Law / Exergy Efficiency of Overall System

Variations of second law / exergy efficiency for various working fluids with evaporator pres-
sure are demonstrated in Figure 5.26. Variations of exergy efficiency follow the similar trend
of thermal efficiency as illustrated in Figure 5.22. Exergy efficiency is calculated as the ratio
of net work output to the net exergy input to the ORC. Exergy input to ORC is almost same
for various refrigerants due to same temperature of GT exhaust. So, working fluid with higher
net work output results in higher exergy efficiency.

However, maximum value of the exergy efficiency reached by ORC is 51%, that means as
much as 49% of the total exergy input is either

carried out by condenser cooling water or


destructed due to irreversibilities
Both of them will be discussed in the following sections. Sequence of refrigerants accord-
ing to exergy efficiency is same as that of thermal efficiency.

60
Figure 5.26 Exergy efficiency variation of selected refrigerants

5.8.2 Exergy carried out by condenser cooling water

Exergy carried out by condenser cooling water is much lower than the exergy destruction, for
example, using refrigerant R1233zd(E), average condenser exergy rejection is approximately
7% of total exergy input. Exergy rejection comparison between various refrigerants is dem-
onstrated in Figure 5.27. R124 results in the highest exergy rejection among all other selected
refrigerants. The sequence of the refrigerants according to condenser exergy rejection from
higher to lower is R124 > R245fa > R1234ze(Z) > R1233zd(E) > R21 > R141b. A close in-
spection of Figure 5.23, Figure 5.26 and Figure 5.27 will reveal that, this sequence is same as
that of condenser heat rejection and just opposite to that of exergy efficiency.

5.8.3 Exergy Destruction of ORC components

Recalling the discussion from the previous section, exergy destruction is a major contributor
to the input exergy consumption. In this section, exergy destruction will be analysed compo-
nent-wise using different refrigerants.

Exergy destruction in the heat exchangers (evaporator and condenser) is primarily due to the
heat transfer irreversibility because of high stream to stream temperature difference.

61
Figure 5.27 Exergy carried out by condenser cooling water for selected refrigerants

For any particular refrigerant, exergy destruction rate in both evaporator and condenser de-
creases gradually with evaporator pressure as demonstrated in Figure 5.28 and Figure 5.29
respectively. This might be due to the fact that as evaporator pressure increases, temperature
difference between the two heat exchanging streams decreases gradually as described in Fig-
ure 5.19. Consequently, there is a reduction in the heat transfer irreversibility which leads to
the decrease in the exergy destruction. However, exergy destruction behaviour for any refrig-
erant is different in evaporator and condenser. For instance, although R21 causes highest
destruction in the evaporator, it results in the lowest exergy destruction in the condenser.
Again, for R141b, the quantity is lower in both the components. From higher to lower, refrig-
erant sequence for evaporator destruction is: R21 > R124 > R1234ze(Z) > R1233zd(E) >
R245fa > R141b and that for condenser destruction: R124 > R245fa > R1234ze(Z) >
R1233zd(E) > R141b > R21. It should be noted that magnitude of exergy rejection values in
the condenser is much larger than those in the evaporator.
Exergy destruction values in the pump are very low compared to the other components. As
shown in Figure 5.30, pump exergy destruction increases almost linearly with evaporator
pressure and entangled lines in the figure indicates that varying the refrigerants in the ORC
does not significantly change the pump exergy destruction values.

62
Figure 5.28 Evaporator exergy destruction for selected refrigerants

Figure 5.29 Condenser exergy destruction for selected refrigerants

As shown in Figure 5.31, turbine exergy destruction values vary significantly as working
fluid is changed in the ORC. Average turbine exergy destruction for R141b and R21 is
approximately double of that for R124. Destruction in the turbine using R1233zd(E),
R1234ze(Z) and R245fa is close to each other.
63
Figure 5.30 Pump exergy destruction for selected refrigerants

Figure 5.31 Turbine exergy destruction for selected refrigerants

64
5.8.4 Exergy Efficiencies of ORC components

According to equation (3.16), for any system or device, higher exergy destruction results in
lower exergy efficiency or vice versa. The exergy efficiencies of the components are demon-
strated in the Figures 5.32-5.34.

Figure 5.32 shows the variation of evaporator exergy efficiency for various refrigerants. It is
evident that, for all the working fluids exergy efficiency of the evaporator increases with the
evaporator pressure. R141b results in the highest exergy efficiency because of the lowest ex-
ergy destruction in the evaporator. As shown in Figure 5.33, turbine exergy efficiency
decreases with evaporator pressure for all refrigerants. Since for turbines exergy destruction
is a function of pressure ratio, so exergy destruction increases with evaporator pressure,
consequently exergy efficiency decreases. Magnitude of condenser exergy efficiency is very
low compared to other components for all working fluids (Figure 5.34). For any refrigerant,
exergy efficiency of the condenser increases with the evaporator pressure because of the
decrease in the exergy destruction values with increasing pressure ratio as explained in the
Figure 5.19.

Figure 5.32 Evaporator exergy efficiency variation for selected refrigerants

65
Figure 5.33 Turbine exergy efficiency variation for selected refrigerants

Figure 5.34 Condenser exergy efficiency variation for selected refrigerants

66
So, for any particular component, a refrigerant that causes higher exergy destruction results in
lower exergy efficiency. Recalling the description in the previous section, the figures of
component exergy efficiency justify this fact. The lower value of the condenser efficiency
results in the lower overall efficiency of the system. Analysing the performance of the ORC
at various evaporator pressure using different refrigerants, it is observed that low GWP fluids
R1233zd(E) and R1234ze(Z) can be used as replacements of R245fa with moderate perform-
ance variation of ORC.

67
6 Conclusions

The major conclusions of the study are:

Thermal efficiency of ORC using selected refrigerants increases with evaporator


pressure. Among the refrigerants considered, R141b results in the highest thermal
efficiency and R124 results in the lowest. This might be attributed to the pressure
ratio for any given refrigerant. R141b exhibits the highest pressure ratio at any
evaporator pressure. For example, at 60 bar evaporator pressure, pressure ratio and
thermal efficiency of ORC for R141b is 45.1 and 19.3% respectively, whereas for
R124, the values are 10.1 and 12.0% respectively.

Exergy efficiency of ORC using different refrigerants exhibit similar behaviour as


the thermal efficiency. R141b results in the highest exergy efficiency (48.5% at 60
bar evaporator pressure) and R124 results in the lowest (30.2% at 60 bar evapora-
tor pressure).

Energy balance indicates that major portion of the input energy is rejected to
condenser cooling water, although exergy balance suggests that its exergy content
is very low. For instance, using R245fa at 10 bar evaporator pressure, heat rejection
is 89% of total energy input but its exergy value is only 7% of total input exergy.

Heat rejection to condenser cooling water decreases with the evaporator pressure.
(for instance, using R245fa, the heat rejection value decreases by 9.1% when
evaporator pressure is increased from 10 to 65 bar). The reason behind these
variations may be attributed to the temperature of the working fluid at the turbine
exit. For all refrigerants, turbine exit temperature decreases with the evaporator
pressure due to greater expansion at the turbine, resulting in lower temperature of
the refrigerant at the turbine exit, consequently lower heat rejection.

A major portion of the input exergy is destructed due to the irreversibility in


various components of the system. Exergy destruction in the heat exchangers
(evaporator and condenser) is higher than other components, primarily due to the
heat transfer irreversibility because of high stream to stream temperature difference
(for instance, at 10 bar evaporator pressure using R245fa, 76% of the input exergy
68
is destructed, condenser itself is responsible for 70% of total exergy destruction).
Exergy destruction rate in both the evaporator and condenser decreases gradually
with evaporator pressure.

It is observed that performance of ORC using two environment-friendly (GWP


value ~ 0) refrigerants, R1233zd(E) and R1234ze(Z) is close to the performance
obtained using R245fa (GWP value ~ 1000) as working fluid. Therefore, with
slight variations of performance, R1233zd(E) and R1234ze(Z) can act as replace-
ments of R245fa which is a commonly used refrigerant in ORC.

Recommendations for future work:

The study is carried out at a particular heat source temperature and flow rate.
Further analysis can be carried out by varying the heat source temperature and
mass flow rates.

More environment friendly refrigerants sustaining the high temperature of the


exhaust stream can be included in the further study.

Exergo-economic analysis can also be conducted further on the existing study


to analyse economic feasibility of the study.

In this thermodynamic analysis, calculations are performed based on simulated


conditions. Appropriate approximations are made to simplify the analysis. It
was not possible to analyse real power systems due to time and resource
constraints. The work can be further extended by taking into account the ac-
tual state conditions.

69
Appendix A
SIMULATION DATA

Performance data of ORC for refrigerant R141b

Table A.1 ORC parameters at different evaporator pressure for R141b

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kg/s) (%) (%)
(kW) (kW)
10 36.05 35 2379 425 11.4 28.8
15 36.2 55 2802 415 13.4 33.9
20 36.35 75 3086 408 14.8 37.4
25 36.51 96 3297 402 15.8 39.9
30 36.68 116 3461 398 16.6 41.9
35 36.85 137 3594 395 17.2 43.5
40 37.02 158 3704 392 17.8 44.9
45 37.19 179 3796 389 18.2 46.0
50 37.37 201 3875 387 18.6 46.9
55 37.56 222 3943 385 19.0 47.8
60 37.74 244 4002 383 19.3 48.5
65 37.93 266 4053 382 19.5 49.1

Table A.2 Exergy Destruction of ORC components for R141b

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
Evaporator Turbine Condenser Pump

10 1609 401 3288 8 582


15 1339 486 3057 13 567
20 1157 549 2896 18 558
25 1024 598 2770 23 551
30 922 639 2667 28 545
35 841 674 2578 32 540
40 775 705 2500 37 536
45 722 733 2431 42 532

70
50 677 759 2367 47 529
55 641 782 2309 52 527
60 611 804 2254 58 524
65 586 824 2203 63 522

Table A.3 Exergy Efficiency variation of ORC components for R141b

Pressure Exergy Efficiency (%)


(bar)
Evaporator Turbine Pump Condenser

10 80.5 85.8 76.2 15.0


15 83.8 85.5 76.3 15.7
20 86.0 85.2 76.3 16.2
25 87.6 85.0 76.3 16.6
30 88.8 84.9 76.3 17.0
35 89.8 84.7 76.3 17.3
40 90.6 84.6 76.4 17.7
45 91.3 84.4 76.4 18.0
50 91.8 84.3 76.4 18.3
55 92.2 84.2 76.4 18.6
60 92.6 84.1 76.4 18.9
65 92.9 84.0 76.4 19.2

71
Performance data of ORC for refrigerant R124

Table A.4 ORC parameters at different evaporator pressure for R124

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kg/s) (%) (%)
(kW) (kW)
10 41.34 17 630 466 3.0 7.6
15 41.44 38 1099 455 5.3 13.3
20 41.53 60 1418 447 6.8 17.2
25 41.63 81 1656 441 7.9 20.1
30 41.72 102 1844 436 8.9 22.3
35 41.82 124 1997 432 9.6 24.2
40 41.91 145 2126 428 10.2 25.8
45 42.01 167 2236 425 10.8 27.1
50 42.1 189 2331 422 11.2 28.3
55 42.2 210 2415 420 11.6 29.3
60 42.29 232 2488 418 12.0 30.2
65 42.39 254 2554 416 12.3 31.0

Table A.5 Exergy Destruction of ORC components for R124

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
Evaporator Turbine Condenser Pump

10 1959 98 4938 4 638


15 1685 175 4675 9 622
20 1498 231 4491 14 611
25 1360 275 4349 19 603
30 1251 311 4234 24 596
35 1163 342 4136 29 591
40 1089 368 4052 34 586
45 1028 392 3977 39 582
50 975 414 3910 44 578
55 929 434 3849 50 575
60 889 452 3793 55 571
65 854 469 3742 60 569

72
Table A.6 Exergy Efficiency variation of ORC components for R124

Pressure Exergy Efficiency (%)


(bar)
Evaporator Turbine Pump Condenser

10 76.3 86.9 76.2 11.4


15 79.6 86.7 76.3 11.8
20 81.9 86.5 76.3 12.0
25 83.5 86.4 76.3 12.2
30 84.9 86.2 76.3 12.3
35 85.9 86.1 76.4 12.5
40 86.8 86.0 76.4 12.6
45 87.6 86.0 76.4 12.8
50 88.2 85.9 76.4 12.9
55 88.7 85.8 76.5 13.0
60 89.2 85.8 76.5 13.1
65 89.6 85.7 76.5 13.2

Performance data of ORC for refrigerant R21

Table A.7 ORC parameters at different evaporator pressure for R21

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kW) (kg/s) (%) (%)
(kW)
10 41.91 30 1911 436 9.1 23.1
15 42.06 51 2482 422 11.9 30.0
20 42.22 72 2863 413 13.7 34.7
25 42.38 93 3145 406 15.1 38.1
30 42.55 115 3365 400 16.1 40.7
35 42.72 137 3542 396 17.0 42.9
40 42.89 159 3690 392 17.7 44.7
45 43.06 181 3814 389 18.3 46.2
50 43.24 203 3921 386 18.9 47.5
55 43.42 226 4013 383 19.3 48.6
60 43.6 248 4094 381 19.7 49.6
65 43.79 271 4165 379 20.1 50.5
73
Table A.8 Exergy Destruction of ORC components for R21

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
Evaporator Turbine Condenser Pump

10 2403 319 3029 7 597


15 2033 432 2728 12 578
20 1780 515 2523 17 565
25 1591 581 2368 22 555
30 1442 636 2242 27 548
35 1321 684 2138 32 541
40 1221 726 2047 38 536
45 1136 764 1967 43 532
50 1064 798 1896 48 528
55 1001 830 1831 53 524
60 947 859 1772 58 521
65 899 887 1717 64 518

Table A.9 Exergy Efficiency variation of ORC components for R21

Pressure Exergy Efficiency (%)


(bar)
Evaporator Turbine Pump Condenser

10 70.9 85.9 76.2 16.5


15 75.4 85.4 76.3 17.5
20 78.5 85.1 76.3 18.3
25 80.7 84.8 76.3 19.0
30 82.5 84.6 76.3 19.6
35 84.0 84.3 76.3 20.2
40 85.2 84.1 76.4 20.8
45 86.2 84.0 76.4 21.3
50 87.1 83.8 76.4 21.8
55 87.9 83.6 76.4 22.3
60 88.5 83.5 76.4 22.7
65 89.1 83.3 76.5 23.2

74
Performance data of ORC for refrigerant R1233zd(E)

Table A.10 ORC parameters at different evaporator pressure for R1233zd(E)

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kg/s) (%) (%)
(kW) (kW)
10 36.85 31 1696 441 8.1 20.5
15 36.99 52 2107 431 10.1 25.5
20 37.13 72 2387 424 11.4 28.9
25 37.28 92 2595 419 12.4 31.4
30 37.42 113 2759 415 13.2 33.4
35 37.57 134 2892 411 13.9 35.0
40 37.72 155 3004 408 14.4 36.4
45 37.87 176 3098 405 14.9 37.5
50 38.02 197 3180 403 15.3 38.5
55 38.17 218 3251 401 15.7 39.4
60 38.33 240 3314 399 16.0 40.2
65 38.49 261 3369 398 16.3 40.8

Table A.11 Exergy Destruction of ORC components for R1233zd(E)

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
Evaporator Turbine Condenser Pump

10 1671 275 4013 7 603


15 1422 351 3783 12 590
20 1254 406 3620 17 580
25 1131 449 3492 22 573
30 1035 485 3388 27 567
35 959 516 3298 32 562
40 897 544 3219 37 558
45 845 568 3148 42 555
50 802 590 3084 46 551
55 766 611 3025 51 549
60 735 630 2971 57 546
65 709 647 2919 62 544

75
Table A.12 Exergy Efficiency variation of ORC components for R1233zd(E)

Pressure Exergy Efficiency (%)


(bar)
Evaporator Turbine Pump Condenser

10 79.8 86.3 76.2 13.1


15 82.8 86.0 76.3 13.5
20 84.8 85.8 76.3 13.8
25 86.3 85.7 76.3 14.1
30 87.5 85.6 76.3 14.3
35 88.4 85.4 76.3 14.6
40 89.1 85.3 76.4 14.8
45 89.8 85.2 76.4 15.0
50 90.3 85.1 76.4 15.2
55 90.7 85.0 76.4 15.4
60 91.1 85.0 76.4 15.5
65 91.4 84.9 76.5 15.7

Performance data of ORC for refrigerant R1234ze(Z)


Table A.13 ORC parameters at different evaporator pressure for R1234ze(Z)

Pressure m ref W pump Wnet m water I II


(bar) (kg/s) (kg/s) (%) (%)
(kW) (kW)
10 33.73 27 1441 447 6.9 17.4
15 33.84 46 1879 436 9.0 22.7
20 33.95 65 2175 429 10.4 26.3
25 34.07 85 2397 423 11.5 29.0
30 34.19 104 2571 419 12.3 31.1
35 34.3 124 2713 415 13.0 32.9
40 34.42 144 2832 412 13.6 34.3
45 34.54 163 2933 409 14.1 35.5
50 34.66 183 3021 407 14.5 36.6
55 34.78 203 3097 404 14.9 37.5
60 34.9 223 3164 402 15.3 38.4
65 35.02 244 3223 401 15.6 39.1

76
Table A.14 Exergy Destruction of ORC components for R1234ze(Z)

Pressure Exergy Destruction (kW) Rejection at Con-


(bar)
denser (kW)
Evaporator Turbine Condenser Pump

10 1785 231 4191 6 612


15 1522 309 3947 11 597
20 1345 365 3775 16 587
25 1213 410 3642 20 579
30 1111 447 3533 25 573
35 1030 478 3440 29 568
40 962 506 3359 34 563
45 906 531 3287 39 560
50 859 553 3221 43 556
55 819 574 3162 48 553
60 784 593 3107 53 551
65 754 610 3056 57 548

Table A.15 Exergy Efficiency variation of ORC components for R1234ze(Z)

Pressure Exergy Efficiency (%)


(bar)
Evaporator Turbine Pump Condenser

10 78.4 86.4 76.2 12.7


15 81.6 86.2 76.3 13.1
20 83.7 86.0 76.3 13.5
25 85.3 85.8 76.3 13.7
30 86.5 85.7 76.3 14.0
35 87.5 85.6 76.4 14.2
40 88.3 85.5 76.4 14.4
45 89.0 85.4 76.4 14.6
50 89.6 85.3 76.4 14.7
55 90.1 85.2 76.4 14.9
60 90.5 85.1 76.5 15.1
65 90.9 85.0 76.5 15.2

77
Appendix B
Pure and Pseudo-Pure fluid properties in CoolProp [81]

Nearly all the fluids modeling in CoolProp are based on Helmholtz energy formulations. This
is a convenient construction of the equation of state because all the thermodynamic properties
of interest can be obtained directly from partial derivatives of the Helmholtz energy.

It should be noted that the EOS are typically valid over the entire range of the fluid, from
subcooled liquid to superheated vapor, to supercritical fluid.

Annoyingly, different authors have selected different sets of nomenclature for the Helmholtz
energy. For consistency, the nomenclature of Lemmon will be used here. Also, some authors
present results on a mole-basis or mass-basis, further complicating comparisons.

Thermodynamic properties of Fluid [81]

In general, the EOS are based on non-dimensional terms d and t , where these terms are de-
fined by,

d = r / rc
t = Tc / T

Where, r c and Tc are the critical density of the fluid if it is a pure fluid. For pseudo-pure mix-

tures, the critical point is typically not used as the reducing state point, and often the
maximum condensing temperature on the saturation curve is used instead.

The non-dimensional Helmholtz energy of the fluid is given by,

a = a0 +ar

Where, a 0 is the ideal-gas contribution to the Helmholtz energy, and a r is the residual Helm-
holtz energy contribution which accounts for non-ideal behavior. For a given set of d and t ,
each of the terms a 0 and a r are known. The exact form of the Helmholtz energy terms is fluid
dependent, but a relatively simple example is that of Nitrogen, which has the ideal-gas Helm-
holtz energy of

a 0 = ln d + a1 ln t + a2 + a3t + a4t -1 + a5t -2 + a6t -3 + a7 ln[1 - exp(- a8t )]

and the non-dimensional residual Helmholtz energy of

78
6 32 36
a r = N kd i t j + N kd i t j exp(-d l ) + N k d i t j exp(-fk (d - 1)2 - b k (t - g k )2 )
k k k k k k k

k =1 k =7 k = 33

and all the terms other than d and t are fluid-dependent correlation parameters.

The other thermodynamic parameters can then be obtained through analytic derivatives of the
Helmholtz energy terms. For instance, the pressure is given by,

a r
p = r RT 1 + d
d t

and the specific internal energy by,

u a 0 a r
= t +
RT t d t d

and the specific enthalpy by,

h a 0 a r a r
= t + +d +1
RT t d t d d t

which can also be written as,

h u p
= +
RT RT r RT

The specific entropy is given by,

s a 0 a r
= t + -a -a
0 r

R t d t d

and the specific heats at constant volume and constant pressure respectively are given by,

cv 2a 0 2a r
= -t 2 2
+ 2
R t d t d
2
a r 2a r
1 + d - dt
c p cv d t dt
= +
R R a r 2 a
2 r
1 + 2 d + d 2
d t d t

79
The EOS is set up with temperature and density as the two independent properties, but often
other inputs are known, most often temperature and pressure because they can be directly
measured. As a result, if the density is desired for a known temperature and pressure, it can be
obtained iteratively. The following algorithm is used to obtain a reasonable guess for the ini-
tial value for the iterative solver:

1. If the fluid is superheated, use a guess of ideal gas ( r = p / ( RT ) )

2. If the fluid is subcooled, use a guess of saturated liquid density

3. If the fluid is supercritical, use a guess of ideal gas ( r = p / ( RT ) )

4. No solution for density as a function of temperature and pressure if the fluid is two-
phase.

80
References

1. Kotas TJ. The exergy method of thermal plant analysis: Elsevier 2013.

2. Haq MZ, Morshed A. Energy and Exergy Based Analyses of a Multi-Fuelled SI engine.
ASME 2013 Power Conference 2013: Paper No. POWER2013-98279.

3. Wark K. Advanced Thermodynamics for Engineers, McGraw-Hill. New York 1995.

4. Moran M. Availability Analysis: A Guide to Efficient Energy Use, American Society of


Mechanical Engineers. New York 1989.

5. Moran MJ, Shapiro HN, Boettner DD, Bailey MB. Fundamentals of engineering
thermodynamics: John Wiley & Sons 2010.

6. Panesar AS. A study of organic Rankine cycle systems with the expansion process performed
by twin screw machines: City University London; 2012.

7. Poullikkas A. An overview of current and future sustainable gas turbine technologies.


Renewable and Sustainable Energy Reviews 2005; 9: 409-43.

8. Ahmadi P, Dincer I, Rosen MA. Exergo-environmental analysis of an integrated organic


Rankine cycle for trigeneration. Energy Conversion and Management 2012; 64: 447-53.

9. Cihan A, Hachafzoglu O, Kahveci K. Energyexergy analysis and modernization


suggestions for a combinedcycle power plant. International Journal of Energy Research
2006; 30: 115-26.

10. Badr O, O'callaghan P, Probert S. Rankine-cycle systems for harnessing power from low-
grade energy sources. Applied Energy 1990; 36: 263-92.

11. Badr O, Probert S, O'callaghan P. Selecting a working fluid for a Rankine-cycle engine.
Applied Energy 1985; 21: 1-42.

12. Rowshanzadeh R. Performance and cost evaluation of Organic Rankine Cycle at different
technologies. 2010.

13. Mols F, Navarro-Esbr J, Peris B, Mota-Babiloni A. Experimental evaluation of HCFO-


1233zd-e as HFC-245fa replacement in an organic rankine cycle system for low temperature
heat sources. Applied Thermal Engineering 2016.

14. Yang M-H, Yeh R-H. Economic performances optimization of an organic Rankine cycle
system with lower global warming potential working fluids in geothermal application.
Renewable Energy 2016; 85: 1201-13.

15. BP (British Petroleum) Statistical Review of World Energy. wwwbpcom.

16. Gibbs JW. The scientific papers of J. Willard Gibbs: Longmans, Green and Company 1906.

17. Rosen MA, Scott DS. Entropy production and exergy destruction: Part Ihierarchy of Earth's
major constituencies. International Journal of Hydrogen Energy 2003; 28: 1307-13.
81
18. Szargut J. International progress in second law analysis. Energy 1980; 5: 709-18.

19. Morris DR, Szargut J. Standard chemical exergy of some elements and compounds on the
planet earth. Energy 1986; 11: 733-55.

20. Haywood R. A critical review of the theorems of thermodynamic availability, with concise
formulations. Journal of Mechanical Engineering Science 1974; 16: 160-73.

21. Sussman M. Mechanochemical availability. Nature 1975.

22. Sussman M. Steady-flow availability and the standard chemical availability. Energy 1980; 5:
793-802.

23. Heberle F, Brggemann D. Exergy based fluid selection for a geothermal Organic Rankine
Cycle for combined heat and power generation. Applied Thermal Engineering 2010; 30:
1326-32.

24. Long R, Bao Y, Huang X, Liu W. Exergy analysis and working fluid selection of organic
Rankine cycle for low grade waste heat recovery. Energy 2014; 73: 475-83.

25. Wang D, Ling X, Peng H. Performance analysis of double organic Rankine cycle for
discontinuous low temperature waste heat recovery. Applied Thermal Engineering 2012; 48:
63-71.

26. Kaka . Energy and exergy analysis of an organic Rankine for power generation from waste
heat recovery in steel industry. Energy Conversion and Management 2014; 77: 108-17.

27. Wu S-Y, Jiang L, Xiao L et al. An investigation on the exergo-economic performance of an


evaporator in ORC recovering low-grade waste heat. International Journal of Green Energy
2012; 9: 780-99.

28. Karellas S, Leontaritis A-D, Panousis G et al. Energetic and exergetic analysis of waste heat
recovery systems in the cement industry. Energy 2013; 58: 147-56.

29. Bronicki L, Harry ZT. Vapor turbines. Google Patents 1962.

30. Li J. Structural Optimization and Experimental Investigation of the Organic Rankine Cycle
for Solar Thermal Power Generation: Springer 2014.

31. Benefits of Triogen. wwwtriogennl/why-triogen/benefits.

32. Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low-temperature organic
Rankine cycles. Energy 2007; 32: 1210-21.

33. Hung T, Wang S, Kuo C et al. A study of organic working fluids on system efficiency of an
ORC using low-grade energy sources. Energy 2010; 35: 1403-11.

34. Lai NA, Wendland M, Fischer J. Working fluids for high-temperature organic Rankine cycles.
Energy 2011; 36: 199-211.

35. Drescher U, Brggemann D. Fluid selection for the Organic Rankine Cycle (ORC) in biomass
power and heat plants. Applied thermal engineering 2007; 27: 223-8.

36. Liu B-T, Chien K-H, Wang C-C. Effect of working fluids on organic Rankine cycle for waste
heat recovery. Energy 2004; 29: 1207-17.

82
37. Maizza V, Maizza A. Unconventional working fluids in organic Rankine-cycles for waste
energy recovery systems. Applied thermal engineering 2001; 21: 381-90.

38. Rayegan R, Tao Y. A procedure to select working fluids for Solar Organic Rankine Cycles
(ORCs). Renewable Energy 2011; 36: 659-70.

39. Angelino G, Gaia M, Macchi E. A review of Italian activity in the field of organic Rankine
cycles. VDI-Berichte 1984: 465-82.

40. Cayer E, Galanis N, Nesreddine H. Parametric study and optimization of a transcritical power
cycle using a low temperature source. Applied Energy 2010; 87: 1349-57.

41. Lecompte S, Huisseune H, van den Broek M et al. Part load based thermo-economic
optimization of the Organic Rankine Cycle (ORC) applied to a combined heat and power
(CHP) system. Applied Energy 2013; 111: 871-81.

42. Zanelli R, Favrat D. Experimental investigation of a hermetic scroll expander-generator.


1994.

43. Wang D, Ling X, Peng H et al. Efficiency and optimal performance evaluation of organic
Rankine cycle for low grade waste heat power generation. Energy 2013; 50: 343-52.

44. Kuo C-R, Hsu S-W, Chang K-H, Wang C-C. Analysis of a 50kW organic Rankine cycle
system. Energy 2011; 36: 5877-85.

45. Hung T-C, Shai T, Wang S. A review of organic Rankine cycles (ORCs) for the recovery of
low-grade waste heat. Energy 1997; 22: 661-7.

46. Tchanche BF, Lambrinos G, Frangoudakis A, Papadakis G. Low-grade heat conversion into
power using organic Rankine cyclesA review of various applications. Renewable and
Sustainable Energy Reviews 2011; 15: 3963-79.

47. Yamamoto T, Furuhata T, Arai N, Mori K. Design and testing of the organic Rankine cycle.
Energy 2001; 26: 239-51.

48. Chen H, Goswami DY, Stefanakos EK. A review of thermodynamic cycles and working fluids
for the conversion of low-grade heat. Renewable and sustainable energy reviews 2010; 14:
3059-67.

49. Tourangeau PR. Montreal Protocol on Substances that Deplete the Ozone Layer: Can It Keep
Us All from Needing Hats, Sunglasses, and Suntan Lotion, The. Hastings Int'l & Comp L Rev
1987; 11: 509.

50. Protocol K. United Nations framework convention on climate change. Kyoto Protocol, Kyoto
1997; 19.

51. Stocker TF. Climate change 2013: the physical science basis: Working Group I contribution to
the Fifth assessment report of the Intergovernmental Panel on Climate Change: Cambridge
University Press 2014.

52. Patten K, Wuebbles D. Atmospheric lifetimes and Ozone Depletion Potentials of trans-1-
chloro-3, 3, 3-trifluoropropylene and trans-1, 2-dichloroethylene in a three-dimensional
model. Atmospheric Chemistry and Physics 2010; 10: 10867-74.

83
53. Myhre G, Shindell D, Bron F et al. Anthropogenic and natural radiative forcing. Climate
change 2013; 423.

54. Kontomaris K. A zero-ODP, low GWP working fluid for high temperature heating and power
generation from low temperature heat: DR-2. The International Symposium on New
Refrigerants and Environmental Technology, Kobe, Japan; 2012; 2012.

55. Hulse RJ, Basu RS, Singh RR, Thomas RH. Physical properties of HCFO-1233zd (E).
Journal of Chemical & Engineering Data 2012; 57: 3581-6.

56. Facts about waste heat. wwwindustrialwasteheatcom.

57. Najjar YS. Efficient use of energy by utilizing gas turbine combined systems. Applied
Thermal Engineering 2001; 21: 407-38.

58. Invernizzi C, Iora P, Silva P. Bottoming micro-Rankine cycles for micro-gas turbines. Applied
thermal engineering 2007; 27: 100-10.

59. Caresana F, Comodi G, Pelagalli L, Vagni S. Micro combined plant with gas turbine and
organic cycle. ASME Turbo Expo 2008: Power for Land, Sea, and Air; 2008: American
Society of Mechanical Engineers; 2008. p. 787-95.

60. Vaja I, Gambarotta A. Internal combustion engine (ICE) bottoming with organic Rankine
cycles (ORCs). Energy 2010; 35: 1084-93.

61. Shu G, Li X, Tian H et al. Alkanes as working fluids for high-temperature exhaust heat
recovery of diesel engine using organic Rankine cycle. Applied Energy 2014; 119: 204-17.

62. Siddiqi MA, Atakan B. Alkanes as fluids in Rankine cycles in comparison to water, benzene
and toluene. Energy 2012; 45: 256-63.

63. Oellrich LR, Leisenheimer B, Srinivasan K. Flammability behavior of (octafluoropropane+


propane) and (octafluoropropane+ methane). Chemie Ingenieur Technik 1997; 69: 984-6.

64. Perrot P. A to Z of Thermodynamics: Oxford University Press on Demand 1998.

65. Rant Z. Exergie, eiu neues Wort fur, technische Arbeitsfahigkeit. Forscliung I ng-Wes 1950;
22: 36-7.

66. Gibbs J. A method of geometrical representation of thermodynamic properties of substances


by means of surfaces: reprinted in Gibbs, Collected Works, ed. WR Longley and RG Van
Name. Transactions of the Connecticut Academy of Arts and Sciences 1931; 2: 382-404.

67. Szargut J. Chemical exergies of the elements. Applied Energy 1989; 32: 269-86.

68. Morshed A. Energy and Exergy Based Analysis of A Multi-Fuelled SI Engine. 2013.

69. Wark K. Advanced thermodynamics for engineers: McGraw-Hill New York 1995.

70. Moran MJ. Availability analysis: a guide to efficient energy use. 1982.

71. Ozdil NFT, Segmen MR, Tantekin A. Thermodynamic analysis of an Organic Rankine Cycle
(ORC) based on industrial data. Applied Thermal Engineering 2015; 91: 43-52.

84
72. Bell IH, Wronski J, Quoilin S, Lemort V. Pure and pseudo-pure fluid thermophysical property
evaluation and the open-source thermophysical property library CoolProp. Industrial &
engineering chemistry research 2014; 53: 2498-508.

73. Goodwin D, Moffat H, Speth R. Cantera: an object-oriented software toolkit for chemical
kinetics, thermodynamics and transport processes, 2009. URL: http://code google
com/p/cantera 2012.

74. Pak P, Suzuki Y. Exergetic evaluation of methods for improving power generation efficiency
of a gas turbine cogeneration system. International Journal of Energy Research 1997; 21:
737-47.

75. Song J, Li Y, Gu C-w, Zhang L. Thermodynamic analysis and performance optimization of an


ORC (Organic Rankine Cycle) system for multi-strand waste heat sources in petroleum
refining industry. Energy 2014; 71: 673-80.

76. Fedele L, Di Nicola G, Brown JS et al. Measurements and correlations of cis-1, 3, 3, 3-


tetrafluoroprop-1-ene (R1234ze (Z)) saturation pressure. International Journal of
Thermophysics 2014; 35: 1-12.

77. KOYAMA S, HIGASHI Y, MIYARA A, AKASAKA R. 3-2. Research and Development of


Low-GWP Refrigerants Suitable for Heat Pump Systems. Risk Assessment of Mildly
Flammable Refrigerants 2013: 29.

78. Fukuda S, Kondou C, Takata N, Koyama S. Low GWP refrigerants R1234ze (E) and R1234ze
(Z) for high temperature heat pumps. International journal of Refrigeration 2014; 40: 161-73.

79. http://wwwpowerfromthesunnet/Book/chapter12/chapter12html.

80. Bejan A, Tsatsaronis G. Thermal design and optimization: John Wiley & Sons 1996.

81. http://wwwcoolproporg/fluid_properties/PurePseudoPurehtml.

85

You might also like