You are on page 1of 57

TECHNO-ECONOMIC POTENTIAL OF WASTE HEAT

RECOVERY WITH ORC TECHNOLOGY USING


DIFFERENT WORKING FLUIDS

Thesis Submitted By

Mamoon Rasheed 2k17-ESE-03

Bashir Ahmad 2k17-ESE-11

Usman Maqbool 2k17-ESE-20

M. Tayyab ul Islam 2k17-ESE-22

Supervisor

Engr. Bilal Bashir (Energy Systems Engineering)

In partial fulfilment of the requirement for degree of B.Sc. Energy Systems


Engineering

DEPARTMENT OF ENERGY SYSTEMS ENGINEERING

NFC INSTITUE OF ENGINEERING & TECHNOLOGY, MULTAN


PAKISTAN

1|Page
In the name of Allah

The most Gracious

The most Merciful

2|Page
NFC INSTITUTE OF ENGINEERING & TECHNOLOGY

TECHNO-ECONOMIC POTENTIAL OF WASTE HEAT RECOVERY


WITH ORC TECHNOLOGY USING DIFFERENT WORKING FLUIDS

This thesis is presented by:

Mamoon Rasheed 2k17-ESE-03

Bashir Ahmad 2k17-ESE-11

Usman Maqbool 2k17-ESE-20

M. Tayyab ul Islam 2k17-ESE-22

under the guidance of their supervisor, and accepted by all the members of the
thesis committee, has been presented to and accepted by the NFC Institute of
Engineering and Technology, Multan in partial fulfilment of the requirements of
the four year degree of B.Sc. Energy Systems Engineering.

. .

Engr. Bilal Bashir

(Supervisor) (External Examiner)

. .

Engr. Mohsin Abdullah Dr. Muhammad Siddique

(Project Coordinator) (Head of Department)

Date .

3|Page
DEDICATION

We dedicate this thesis to our Parents who has supported us a lot in every step of life, to our
Teachers who guided us and make us eligible to set higher targets & to our Supervisor Engr.
Awais Ahmad & Engr. Bilal Bashir who has been a constant source of knowledge, inspiration
and motivation for us.

4|Page
ACKNOWLEDGEMENT

First of all, thank to Allah Almighty and the greatest reformer Hazrat Muhammad (peace be
upon him) with the help of their blessings, we are able to complete our project and graduation.
Secondly we are thankful to our parents for their continuous support.

We show our deepest gratitude to our supervisor Engr. Bilal Bashir & Engr. Awais Ahmad for
giving us an opportunity of designing a project entitled “TECHNO-ECONOMIC POTENTIAL
OF WASTE HEAT RECOVERY WITH ORC TECHNOLOGY USING DIFFERENT
WORKING FLUIDS” and guiding us till the end of this project.
We are also thankful for entire Energy Systems faculty for being a source of inspiration and
information for us.

I would also like to thank my friends, my class fellows from Energy System Engineering
Department they help us in project and encourage throughout the project.

5|Page
ABSTRACT
Renewable energy resources are becoming popular alternative due to its environment friendly
applications and on the other hand, energy efficient technologies are also becoming an
important contributor in reducing the greenhouse gas emissions because less energy is
consumed for providing the same amount of energy supply, that ultimately decreasing the
overall emissions and hence improved sustainability of the system. The potential for recovering
heat from industrial waste heat sources has been particularly promising. Industries like
petrochemical, steel, cement, petroleum and some others are high density energy consumers as
a result their waste heat recovery potential is also significant and for this reasons, several hybrid
systems have gained popularity to reduce the overall waste heat consumption on national and
international level.

Organic Rankine (ORC) cycle-based systems became popular in the last 2 decades for power
generating systems through heat in various applications. In comparison with the ordinary
Rankine cycle, the ORC-based power systems have freedom to choose working fluids and
expansion machines, as an additional degree of flexibility, allowing optimal configurations
regarding the thermodynamic as well as techno-economic aspects. ORC systems can be
designed to optimally convert waste heat from different sources like internal combustion
engine exhaust, geothermal heat sources, biomass applications, solar thermal applications, and
other thermal gradient-based sources like ocean thermal energy conversion (OTEC). It has
several perks to use like the maintenance-free automated operation and relatively smaller
installation and operational costs compared to Rankine cycles which make them ideal for
commercial use specifically in <10MW projects.

This study is about the technical potential of the waste heat recovery from the Organic Rankine
Cycle. It involves the waste heat potential, designing of the ORC system and the comparison
of the performance of two different working fluids for the ORC system.

It focused on the investigation of a small scale organic Rankine cycle system which utilize low
grade heat from waste steam for a power generation application. Steam is used in a plate-type
heat exchanger and it condenses, while the working fluid evaporates at the same time from the
condensing heat of the steam. Typically, experimental organic Rankine cycle systems used
intermediate heat transfer fluids such as thermal oil or pressurized hot water to be first heated
by a heat source and then the intermediate fluid will circulate in the ORC system evaporator
for waste heat recovery applications. Simultaneous steam condensation and working fluid

6|Page
evaporation in the same heat exchanger is unique in small scale size range for ORC application
to observe in this size range. The reduction of the intermediate loop allows a heat transfer loop
to be reduced which reduces the cost and the greater temperature of the working fluid
evaporation can be obtained to increase the thermal efficiency. This work offers the overall
performance of the cycle and then the performance of the individual components (expanders,
pumps, etc.) for a small scale ORC system in the small scale size range.

The total efficiency of an ORC system using R245fa is obtained as 19.88 while efficiency of
turbine is 20.99%. On the other hand, R123 gives the total system efficiency of 17.6% while
efficiency of turbine is 21.06%.

7|Page
TABLE OF CONTENTS

1. Introduction

2. Literature Survey
2.1. Sources of Heat
2.2. Waste Heat Recovery
2.3. Development of a waste heat recovery system
2.4. Waste heat recovery in cement plants
2.5. Power availability in cement plants

3. Organic Rankine Cycle


3.1. ORC technology & applications
3.2. Biomass combined Heat & Power
3.3. Geothermal Energy
3.4. Solar Power Plants
3.5. Waste Heat Recovery
3.6. Heat Recovery on internal combustion engines
3.7. ORC manufacturers & market evolution
3.8. Comparison with the steam Rankine cycle
3.9. Working fluid selection

4. Data Collection & Design


4.1. Equipment Details

5. Methodology & Calculations


5.1. R-245fa
5.2. R123

6. Conclusion

7. References

8|Page
LIST OF FIGURES

Figure Description

Fig 1 Composite curve of an industrial process

Fig 2 Diagram of an evacuated tube solar collector

Fig 3 Potential heat to recover in a cement plant

Fig 4 WHR from kiln exhaust gas for preheating combustion air around 1906

Fig 5 Plant Layout for Organic Rankine Cycle

Fig 6 Proposed ORC system cycle

Fig 7 T-S diagram of proposed ORC system cycle

Fig 8 ORC system using R245fa

Fig 9 ORC system using R123

9|Page
CHAPTER 1- INTRODUCTION

A global trend for sustainable and efficient utilization of available energy sources is growing
because of the global warming and continuous depletion of conventional fossil fuel resources.
The International Energy Outlook 2013 (IEO2013) has forecasted that world energy
consumption will grow by 56% till 2040i. The existing energy resources (natural gas,
petroleum, coal, and uranium) are depleting rapidly and a time will reach when they will vanish
completely for future generations. The persistent and widespread dependence on fossil fuel
resources has become a major reason of their depletion. It has also challenged environment by
increasing air pollution and greenhouse gases. These greenhouse gases have become a major
reason for the global warming. In addition, burning of fossil fuels has added approximately
21300 million tons of toxic carbon dioxide (CO2) every year. It is reported that only half of this
amount is absorbed by natural processes, this means there is net increase of 1065 million tons
of atmospheric CO2 every year.

Renewable energy resources are becoming popular alternative due to its environment friendly
applications and on the other hand, energy efficient technologies are also becoming an
important contributor in reducing the greenhouse gas emissions because less energy is
consumed for providing the same amount of energy supply, that ultimately decreasing the
overall emissions and hence improved sustainability of the system. The potential for recovering
heat from industrial waste heat sources has been particularly promising. Industries like
petrochemical, steel, cement, petroleum and some others are high density energy consumers as
a result their waste heat recovery potential is also significant and for this reasons, several hybrid
systems have gained popularity to reduce the overall waste heat consumption on national and
international level.

Organic Rankine (ORC) cycle-based systems became popular in the last 2 decades for power
generating systems through heat in various applications. In comparison with the ordinary
Rankine cycle, the ORC-based power systems have freedom to choose working fluids and
expansion machines, as an additional degree of flexibility, allowing optimal configurations
regarding the thermodynamic as well as techno-economic aspects. ORC systems can be
designed to optimally convert waste heat from different sources like internal combustion
engine exhaust, geothermal heat sources, biomass applications, solar thermal applications, and
other thermal gradient-based sources like ocean thermal energy conversion (OTEC). It has
several perks to use like the maintenance-free automated operation and relatively smaller

10 | P a g e
installation and operational costs compared to Rankine cycles which make them ideal for
commercial use specifically in <10MW projects.

This study is about the technical potential of the waste heat recovery from the Organic Rankine
Cycle. It involves the waste heat potential, designing of the ORC system and the comparison
of the performance of two different working fluids for the ORC system.

It focused on the investigation of a small scale organic Rankine cycle system which utilize low
grade heat from waste steam for a power generation application. Steam is used in a plate-type
heat exchanger and it condenses, while the working fluid evaporates at the same time from the
condensing heat of the steam. Typically, experimental organic Rankine cycle systems used
intermediate heat transfer fluids such as thermal oil or pressurized hot water to be first heated
by a heat source and then the intermediate fluid will circulate in the ORC system evaporator
for waste heat recovery applications. Simultaneous steam condensation and working fluid
evaporation in the same heat exchanger is unique in small scale size range for ORC application
to observe in this size range. The reduction of the intermediate loop allows a heat transfer loop
to be reduced which reduces the cost and the greater temperature of the working fluid
evaporation can be obtained to increase the thermal efficiency. This work offers the overall
performance of the cycle and then the performance of the individual components (expanders,
pumps, etc.) for a small scale ORC system in the small scale size range.

11 | P a g e
CHAPTER 2- LITERATURE SURVEY

Energy has played a key role in shaping the world as we see it today. Fossil fuels make up most
of the world's energy consumption, but their estimates are limited because the rate of resource
utilization is much higher than the rate of discovery of new reservoirs. Due to the rapid
depletion of fossil fuel resources and their environmental impacts, the world is being
challenged to shift its energy dependence from fossil fuels to renewable energy sources and
efficient use of available energy sources.

Global energy consumption has reached an unpredictable level, releasing large amounts of CO2
into the atmosphere. Current concerns about climate change call for measures to reduce
greenhouse gas emissions, most likely including the following modifications to existing energy
systems.

i. A decrease in the energy intensity of buildings and industry.


ii. A shift from fossil fuels toward electricity, e.g. for transportation and space heating.
iii. Clean power generation by a massive shift toward renewable energies, comprising wind
energy, PV, CSP, biomass, geothermal and large hydro.
iv. Grid capacity and strengthening of inter-regional transmission lines to absorb daily and
seasonal fluctuations.

In order to reduce carbon dioxide emissions from the energy and transportation sectors, while
still providing reliable and affordable service, innovation is required in the areas of power
generation and energy efficiency. There are a wide variety of low-temperature heat sources,
such as heat lost from industry and transportation, solar thermal, biomass and geothermal,
which contain large amounts of energy, but use traditional power generation techniques.
However, the temperature is not high enough to be economically viable.

The steady increase in energy demand is raising questions about how the energy production
network should be developed in the coming years. Recent policy decisionsii have identified
reduction of emissions and improvement of the efficiency of the conversion system as energy
targets. To this end, experts agree that one of the key approaches to achieving this goal is power
generation at the site to minimize the losses and improve efficiencyiii. It is based on the
development of several small power plants that are used directly for power generation for
consume in the vicinity of production sites, avoiding transmission losses and reducing carbon
emissions. In addition, the use of wasted heat to achieve the goals set by government regulators

12 | P a g e
is important. In the EU, 800 TWh of thermal energy is released into the atmosphere each year
as wasted heat from industrial processesiv. The scenario described has led the scientific
community to explore new technologies capable of generating efficient local electricity.

3.2. Sources of Heat:

Following are the common various heat sources available for heat:

Waste Heat:

Industrial processes such as blast furnaces, ceramic kilns, chemical reactors, as well as large
amounts of waste heat reject heat that can no longer be used for this processv. This represents
a significant financial loss for the industry as well as increasing emissions from both chemical
and thermal pollution. Waste heat is responsible for approximately 4.4% of the installed ORC
capacity globally, primarily in small-scale applicationsvi. Of these, metal production, cement
and lime, and glass manufacturing are the most important.

The UK Department of Business Energy and Industrial Strategyvii report that industrial waste
heat represents a total of 47.9TWh/year, or an average power of 5.46GW. Of this resource,
29.6% exists at high temperature (>500°C), 5% exists at medium temperature (250-500°C),
and the remaining 65.4% exists at low temperature, or lower than 250°C. Cayer et alviii report
that in Canada, two thirds of primary energy consumption by the eight largest industrial sectors
is released as waste heat. Assuming a 10% utilisation rate, this implies a potential annual
recovery of 168PJ (an average power of 5.3GW), and reduction in Greenhouse Gas emissions
of 11.2MT. Hung et alix report that in Taiwan, 88% of waste heat sources are medium to low
temperature heat sources, existing at lower than 370°C. In a different paper they state that low
grade waste heat accounts for 50% of total heat generated in industry, leading to thermal
pollution. Budisulistyo et al. estimate that the between 20 and 50% of energy from industry is
released as waste heat.

Wang et al. point out that waste heat sources do not always form straight lines on a T-H
diagram. They can consist of several different waste heat streams, each with their own thermal
characteristics, which are normally combined to produce a composite curve, as shown in Fig1.
They classified these composite curves into straight, concave and convex curves, and compared
ORCs and Kalina cycles being powered by heat sources of each type.

13 | P a g e
Fig 1: Composite curve of an industrial process

Solar Power:

Solar power is a special application for ORCs, in which the global installed ORC capacity is
less than 1 [4]. However, there are several ways to concentrate solar energy to generate a
temperature useful for the organic raincan cycle. An emptied tube consists of solar collector
glass tubes with a vacuum. Run secondary tubes through these tubes that contain working fluids
and are coated with absorbent material, as shown in Fig 2 [13].

Fig 2: Diagram of an evacuated tube solar collector

14 | P a g e
The temperature of the thermal fluid supplied by the solar collector will vary throughout the
year. Zhang et al. reported a liquid temperature of 217.4 ° C in summer and 137.0 C in winter
in Osaka, Japan, with an average annual value of 177.4 C. This change in temperature can
cause significant problems for the cycle. If complete evaporation is to be ensured during the
winter months, the cycle must be run with either a variable extension ratio, or with a significant
superheat during the summer months. Both of these introduce technical challenges to the cycle.
Woo et al. found that they could produce a thermal fluid temperature of 95 ° C using flat-plate
solar collectors, and that the total cost of the cycle was dominated by solar collectors. Solar-
powered ORC system Zhang et al. introduced a 200-kilowatt ORC power plant using parabolic
trough collectors in Tianjin, China. Finds the performance of the collector of 57 of, and the
performance of the first law of 15 of using the maximum cycle temperature of 145 °C.

In addition, although high heat source temperature increases the efficiency of the ORC, they
reduce the efficiency of the solar collector, heat loss from the collector to the atmosphere also
increases with increasing thermal fluid temperature. Helusi et al. observed a change in collector
performance from a temperature range of 40 ° C to 95 ° C to 60% to 45%. This introduces
corrective conflict in the cycle. The temperature of the thermal fluid supplied to the ORC by
the solar collector is strongly affected by the area of the solar collector. High temperatures
increase the cycle's power generation, but also reduce its efficiency in the use of energy in
thermal fluids, while increasing costs.

Geothermal

Geothermal heat sources are a common heat source considered for the application of the
Organic Rankine Cycle, with a range of heat source temperatures and cycle configurations
appearing in the literature. Tartière et al. report that 74% of global installed ORC capacity is
geothermal, although in terms of number of plants this share is smaller, with geothermal
systems tending to have a higher power output than other applications. Barbier estimates that
around 70% of the global geothermal energy potential exists at temperatures lower than 130°C.

Binary Geothermal Cycles use an ORC, drawing the heat for its evaporator from the primary
geothermal fluid. This differs from a single flash cycle, in which the working fluid for the cycle
is the geothermal fluid itself. Binary cycles are generally more efficient at lower heat source
temperatures, as well as generally having longer plant lifetimes due to the fact that the highly
aggressive geothermal brine is kept physically separate from the moving parts of the ORC. A
large amount of research has been carried out on the application of the ORC to geothermal heat

15 | P a g e
sources. Walraven et al. analysed geothermal powered ORCs with the aim of comparing water-
and air-cooled systems. They point out that the parasitic power of air cooled systems can be
twice that of comparable liquid cooling systems, and the capital cost of components 150%
those of their equivalents, however, this extra expenditure can be justified in regions where
sufficient cooling water is not readily available. They took the details of a geothermal project
in Belgium, giving them a heat source temperature of 125°C. Liu et al. investigated various
typical geothermal heat source temperatures from 80°C to 180°C, while using R245fa as the
working fluid. Gu et al. estimate the parasitic power in geothermal cycles, used to pump thermal
fluid, coolant, fans and other auxiliaries, to be about 15% of the total power output.

Bottoming Cycles

ORCs are often used as bottoming cycles for higher-temperature cycles, such as diesel engines,
Brayton Cycles or other Rankine Cycles. The systems can be scaled to systems varying in size
from road vehicles to large ships or land based power stations. Roughly 8% of global installed
ORC capacity is in the form of bottoming cycles, making them the third most widespread
application for the ORC.

As with all thermal cycles, these cycles necessarily reject large amounts of heat. Diesel engines
reject about 27% of their fuel energy through the exhaust, 20% through the coolant, and 7%
through the intercooler. Typical resource temperatures range from 160°C to 380°C, with a large
variation in between these values. The effects of using ORCs to improve the performance of
other thermal cycles has been investigated in the literature. Chen et al. reported the use of a
bottoming cascade-type ORC capable of increasing the peak thermal efficiency of a diesel
engine from 45.3% to 49.5%. Michos et al. reported a 9.1-10.2% net reduction in brake specific
fuel consumption for a V12 marine diesel engine when a bottoming ORC was added. Eveloy
et al. recorded an increase in first law efficiency of 12% compared to a gas turbine operating
without a bottoming cycle. When using ORCs as bottoming cycles, attention must be paid to
the issue of back-pressure, which can reduce the efficiency of the topping cycle, reducing the
benefit provided by the bottoming cycle or even cancelling it out completely. Additionally,
some topping cycles, such as road vehicle engines, will result in a heat input that varies over
time, which can result in the bottoming cycle operating at part-load, negatively affecting its
own performance.

16 | P a g e
3.2. WASTE HEAT RECOVERY

In the present era, the major issue that all types of firms and organizations face is their
increasing energy consumption and they want to decrease their energy consumption, ultimately
cost and environmental footprint. This can be achieved by various methods and one of them is
energy production from environmental friendly sources, known as Waste Heat Recovery.

Waste Heat:

Amount of heat that is generated during the operation process of the plant, by either fuel
combustion or chemical reaction, and exhausted into the atmosphere but it can still be
recovered for some useful and economic purposes.

The strategy for how to restore this heat depends in part on the temperature of the waste heat
gases and the economics involved. Boilers, kilns, ovens and kilns produce large amounts of hot
gas. If some of this lost heat could be recovered, much of the basic fuel used to generate
electricity could be saved. The energy lost in waste gases cannot be fully restored. However,
most of the heat can be used for advantage in different ways and the losses can be minimized.

In the last years, large endeavours have been made to extend the market share of renewable
energies. Power and heat cogeneration by waste heat is one of the most interesting options for
a sustainable and reliable energy supply due to its high availability. The possibility of
recovering heat from engine paths and industrial waste heat sources has been particularly
promisingx. Waste heat sources are according to the temperature rangexi:

➢ High temperature (> 650 C),


➢ Medium temperature (230-650 C),
➢ Low temperature or low temperature (<230 C).

Analysis of statistics shows that low-grade waste heat accounts for 50% or more of the total
heat generated in the industryxii.

Heat recovery

In any heat recovery application, it is important to know how much heat can be recovered and
also how this energy source can be used. The equation for calculating the total recoverable heat
is give in following eq. (i)

𝑄 = 𝑚°. 𝐶𝑝 . ∆𝑇……….(i)

17 | P a g e
where,

𝑄 = amount of heat content (kJ/s)

𝑚° = mass flowrate of the substance (kg/s)

𝐶𝑝 = specific heat of the substance (kJ/kg ˚C)

∆𝑇 = temperature difference (˚C)

Since 𝑚° = 𝜌. 𝑉°

𝜌 = density of the flue gas (kg/m3)

𝑉°= flowrate of the substance (m3/s)

Waste heat can be used either for other processes that require thermal energy, heating or drying
purposes, or for electric power generation. Subsequent use as heat for thermal processes or
heating, which is usually more energy efficient than power generation, requires either on-site
heating users or near the facility, or nearby district heating networks. However, this is not
usually the case, especially since seasonal heat supply and demand vary. If such heat is of no
use, generating electricity from waste heat may be an option. Electricity can either partially
meet the power demand of the plant or it can be fed to the public power grid.

3.2. Development of a waste heat recovery system

Understanding the process is essential for the development of a waste heat recovery (WHR)
system. This can be completed by reviewing flow sheets, layout diagrams, piping isometrics,
electrical and instrumentation cable ducting, etc. Necessary factors that must be considered are:

❖ Sources and uses of waste heat


❖ Upset conditions occurring in the plant due to heat recovery
❖ Availability of space
❖ Any other constraint, such as dew point occurring in any equipment etc.

After identifying waste heat sources and their potential use, the next step is to select the
appropriate heat recovery systems and equipment to recover and use. It is important to review
selected waste heat recovery systems based on financial analysis such as investment,
depreciation, payment term, rate of return, etc. In addition, it is important to seek the advice of
experienced consultants and suppliers for rational decision making.

18 | P a g e
3.2. Waste heat recovery in cement plants

Cement plants are huge consumers of energy because their heat dissipation is 3.0 - 3.5 GJ/t
clinker and power consumption is approximately 100 to 100 kWh/t (cement in bin). Heat is
generated largely from fossil fuels and in some countries partly from alternative fuels, in most
cases from well-defined waste segments.

Most of the electricity used is generated in power plants that are mostly run on fossil fuels.
Therefore, the energy consumption of the cement plant produces direct and indirect CO2
emissions. This is a very important factor to consider when inspecting the installation of a
WHR system as the needs of industries to reduce greenhouse gas emissions increase year by
year.

Increasing energy efficiency for environmental and economic reasons is one of the current
business needs. Therefore, cement plants need to strengthen their efforts to increase energy
efficiency and therefore reduce energy consumption.

The first step is to reduce power consumption. The cement plant has a lot of potential to
increase power consumption efficiency. The reduction in power consumption is the result of a
thorough analysis of the entire cement production process. As a result of the analysis, many
(relatively small) steps will be realized with a large number of power consumers.

Second, from the point of view of energy and climate change, the most important process is to
reduce fossil fuel consumption. The basic idea is to move the plants as much as possible under
"adaptive operating conditions" (during this process the system is thermodynamically isolated
- there is no heat transfer to the surroundings except for wall damage). In other words, the waste
heat will be restored to the maximum. This can be achieved through very few important steps.
After economically increasing the efficiency of the cement plant, the remaining heat can be
converted into electricity.

3.2. Power Availability in Cement Plants

Waste heat from cement manufacturing can be obtained by either one or two sources which are
as follows:

➢ From the clinker air, or


➢ From the waste gas after the preheater tower.

19 | P a g e
Fig 3: Potential heat to recover in a cement plant

The power plant in accordance with cement making process is an installation of secondary
importance. Therefore, the introduction of power plant must not disturb the process of cement
manufacturing. This technique is accomplished by installing the heat exchangers in a bypass
mode as shown in fig 3. Under normal operating conditions, the waste heat and the air cooling
the clinker flows through the heat exchangers. If for any reason the power plant should be out
of operation, while the kiln is operating at its full capacity, the waste gas bypasses the heat
exchanger and flows through the alternate route to the cooling tower, and the clinker cooler air
flows through the air cooler before it enters the dust precipitator.

With the introduction of rotary kilns in the manufacturing of cement clinker later in 19th
century, clinker was used to be burned in wet and dry rotary kilns without installing preheater
that is very well known technology now a days. The exhaust gas after leaving the kiln furnace
has temperature ranges between 600 and 1000°C according to the type and length of the kiln.
Lost heat was recovered for the drying of the raw materials as well as for the preheating of the
combustion air. Surprisingly, in recent times, even steam boilers were used for waste heat
utilization in the starting era of cement production. The fig 2 below shows the utilization of
heat in flue gas by a system of heating tubes as it was performed more than a hundred years
ago. As lost heat in flue gas exceeded the heat required for raw material drying and combustion
air preheating, lost flue gas heat was already being recovered by steam boilers in the first
decades of the last century.

20 | P a g e
Fig 4: WHR from kiln exhaust gas for preheating combustion air around 1906

In the last few years, waste heat recovery systems experienced a revival. Today, in the Middle
and Far East more than 100 applications in the cement industry are known, in the beginning
initiated especially in cases with high energy process or with a low service security of the public
electric power grid. State-of-the-art technology for waste heat recovery systems, the
development of energy prices, and the growing need to improve energy efficiency all provide
new opportunities for implementing waste heat recovery systems in the already highly efficient
kiln systems of the cement industry.

21 | P a g e
CHAPTER 3- ORGANIC RANKINE CYCLE

The Organic Ranking Cycle (ORC) is accepted as a viable technology for converting low
temperature heat into electrical power due to organic operation, low maintenance, and
favourable operating pressurexiii. The use of ORC system in waste heat recovery is also
beneficial as there is no interference in the main work of the operator and environment friendly
operation as no emissions are involved in the environmentxiv. Furthermore, the installation of
ORC-based systems helps reduce greenhouse gas emissions, which would be generated if fossil
fuels were used to generate the same electricityxv.

ORC power systems generally consist of four major components which are; evaporator,
condenser, pump, and expansion valve. The evaporator and condenser are basically heat
exchangers, sized accordingly for a certain duty to operate with specified fluids at specific
operating conditions. The current state of the art can be considered sufficient enough to support
the technological requirements to ensure the availability of heat exchangers to be used as
evaporators and condensers. The pumps have also been well developed and can be bought off
the shelf to fulfil the requirements of the ORC system. However, the expander can be
considered as the most technically advanced component of an ORC system. The expander or
the expansion valve is the machine which extracts the energy from the expansion of high-
pressure vapor resulting in low pressure while passing through its inlet to the outlet port and
converts fluid energy to mechanical power which can be rotational or reciprocating, and is then
often converted to electrical power via direct or indirect coupling to a generator. Organic
Rankine cycles have relatively low thermodynamic efficiency due to limited temperature
differences between the heat source and heat sink streams. Therefore, the efficiency of the
overall cycle is highly sensitive to the efficiency of the expansion machine Ill. Therefore, the
selection of an appropriate expander for a certain ORC application is of great importance to
avoid further efficiency reductions and for commercial viability. Depending on the application,
operating conditions (temperature, pressure, and mass flow rate), working fluid, and power
levels, different types of expansion machines can be used.

The Organic Ranking Cycle (ORC) is widely regarded as a promising technology that produces
electrical energy from low-grade thermal sources. Many power plants around the world have
been installed in the megawatt range in the last decade. However, despite its market potential,
the commercialization of ORC power plants in the kW range has not reached a high level of
maturity for a number of reasons. First, the specific cost to offer an attractive payback period

22 | P a g e
is still too high, and second, potential buyers of small-scale ORCs are usually SMEs (small to
medium-sized enterprises), usually less aware of savings potential of this technology. When it
comes to small-scale plants, additional design issues arise that still limit the widespread
availability of technology.

The Organic Ranking Cycle (ORC) gained attention because it is capable of generating
electrical energy in combination with a renewable energy source and therefore converts the lost
thermal energy into electricity. These capabilities make this technology suitable for a variety
of applications, such as biomass, geothermal, solar and waste heat recovery.

Recent studies on heat recovery in energy-related industries have shown that, in the simplest
scenario, approximately 20,000 GWh of thermal energy can be saved and 7.6 M tons of CO2
can be minimized from the aid of the ORC technology in the EU industrial sectorxvi. Despite
the fact that ORC technology is a promising solution for waste heat recovery, commercial small
scale units are few and far between and the technology has yet to reach a high level of
readinessxvii.

Electrical power is usually generated in a process based on the Rankine cycle with water, which
is the working fluid. ORC-based power systems are similar to traditional Rankine cycle
systems, but use organic compounds as working fluids instead of water. Unlike water, turbine
expansion extends to most organic fluids that are not in the wet steam regime but above the
condenser temperature in the gas phase. Thus, internal heat exchangers are often used to
improve efficiency as shown in fig 3. In-depth investigations are underway into ORC-based
systems for waste heat recovery, biomass applications, geothermal heat recovery and solar
thermal applications.

Fig 5: Plant layout for Organic Rankine Cycle

23 | P a g e
Compared to water, organic fluids are beneficial when the maximum temperature is low and /
or the power plant is small. At lower temperatures, organic fluids result in higher cycle
performance than water. In small plants, organic fluids are preferred, as fluid mechanics leads
to higher turbine efficiency even at partial loadsxviii. This is the main reason for using ORC for
low grade waste heat application. Another benefit of ORC in small plants is the legal and
economic benefits. Water works well at high pressures, which requires increased safety
measures that are not economically viable for small scale plants.

ORC is not a new concept and much research has been donexix. The research was mainly
focused on low temperatures. Typical applications use geothermalxx or waste heatxxixxii. It has
also been tested as a bottom cycle in conjunction with gas turbinesxxiii or other high temperature
cycles.

The choice of working fluid is one of the key decisions for maximum efficiency of ORC based
power system in terms of efficiency, net power generation and environmental impactxxiv.
Working fluids can be characterized as wet, entropic and dry fluids. Wet fluids are considered
unsuitable for ORC application because droplets can form in the expansion machine and
adversely affect the life and performance of the expansion machine. The dry fluid does not
emit any droplets but occurs in an extremely hot state after the expansion of the turbine, the
energy contained in this heat vapor needs to be recovered using a recuperator or released as a
condenser load. Isotropic fluids result in a balanced average of dry and wet fluid types but are
not preferred due to the chemical instability, cost, and environmental effects that leave dry
fluids as the best choicexxv. Following are a few general parameters that must be consider in
order to select the optimum working fluid for organic Rankine cycle:

❖ Thermodynamic properties,
❖ Fluid stability and its compatibility with materials in contact,
❖ Safety, health & environmental aspects, &
❖ Availability & Cost

Walraven et al. suggests that the choice of suitable working fluid depends strictly on the
conditions (heat source and sink temperature)xxvi. Leo et al.xxvii suggests that the performance
of the ORC system is a weak function of the vapor temperature of the system and the critical
temperature of the selected working fluid. Li et al also saw a link between the critical
temperature of the working fluid and the economic operationxxviii. Xu et al. suggested that the
thermal utilities are well associated with critical temperatures and that R245fa can be used over

24 | P a g e
a wide range of heat temperaturesxxix. Wang et al. suggested that R245fa was the best choice
among R245fa, R600a and R134axxx for the source temperature of 89.85 C.

Based on the overall performance, thermodynamic optimization of the waste heating system,
it is recommended that n-butane, R245fa, R1123, n-pentane, HFE7000, SES36 and R134a and
R1234yf are candidates in order to reduce the overall efficiencyxxxi. R236ea, R245ca, R245fa,
R600, R600a, R601a, RE134 and RE245 are suitable options for low temperature organic
ranking cycle systems based on the use of state equations for thermodynamic screeningxxxii.

The choice of expansion machine is another important parameter that affects the performance
and reliability of ORC based power systems. Depending on the size of the expander type
selection system and the selected working fluid. Not all types of expansion machines can be
matched for imposed working conditions because of working fluid pressure and flow rate
valuesxxxiii. Different types of machines have their own advantages and disadvantages as
presented in Table 1.

Table 1: The comparison of various types of expander suitable for ORC system.

Capacity
Rotate speed
Type range Cost Advantages Disadvantages
(rpm)
(kW)
High cost, low
Light weight,
efficiency in
mature
Radial-inflow off-design
50–500 8000–80,000 High manufacturability
turbine conditions and
and high
cannot bear
efficiency
two-phase
High efficiency,
simple
Low capacity,
manufacture,
Scroll lubrication and
1–10 <6000 Low light weight, low
expander modification
rotate speed and
requirement
tolerable two-
phase
Tolerable two- Lubrication
phase, low rotate requirement,
Screw
15–200 <6000 Medium speed and high difficult
expander
efficiency in off- manufacture
design conditions and seal
Reciprocating High pressure Many
piston 20–100 – Medium ratio, mature movement
expander manufacturability parts, heavy

25 | P a g e
, adaptable in weight, have
variable working valves and
condition and torque impulse
tolerable two-
phase
Tolerable two-
phase, torque Lubrication
Rotary vane
1–10 <6000 Low stable, simple requirement and
expander
structure, low low capacity
cost and noise

Qiu et al.xxxiv concluded that 1-10 kW power output is a good choice for ORC based micro
CHP systems within scroll capacity. Multi-van expanders are a cheap and mechanically easy
alternative to other extension devices for small-scale ORC systems if sealing issues are
resolvedxxxv. Lemort et al modified an air compressor to test as a working fluid with R123 in
the ORC application. Based on the results, a semi-experimental simulation model of the
expander was developed that can accurately predict the mass flow rate, shaft power and exhaust
temperature. The device can achieve an isentropic effect of 68%xxxvi. Retrofitting of
refrigeration compressor with R245faxxxvii achieved 71.03% isentropic effect for nominal
electrical output power of 2.5 kW. Chang et al. The 2.3 kW scroll expander offered a maximum
expansion anthropic performance of 73% with a built-in volume ratio of 4.05 which was
originally an open drive scroll air compressorxxxviii.

Declaye et al. showed the isentropic efficiency of 75.7% for 2.1 kW power generation to
facilitate ORC system design due to the absence of oil-free expander oil recovery
componentsxxxix. In general, riffled scroll machines had low efficiency due to suction /
discharge losses, mechanical losses, over and under expansion losses and leakage losses.
Various measures were taken to improve the performance of the expander, such as sealing,
smoothing and modification of scroll geometryxl. Therefore, a scroll machine designed
specifically for the ORC-based power system by the Air Squarexli has been adapted as the best-
fitting expansion machine.

ORC experimental investigations have been conducted for different applications and sizes. Pei
et al.xlii reported the performance of a 3.75 kW scale ORC system, using R123 as a working
fluid, custom designed turbine, which can achieve 0.65% isentropic efficiency, 6.8% cycle
efficiency, at a temperature difference of 70℃, with about 1 kW of shaft power, thermal oil
was used by artificial heating. The feasibility of low temperature ORC application using R123

26 | P a g e
working fluid is demonstrated experimentally with a turbine customized by Li et al with a
source temperature of 80℃ to achieve a thermal efficiency of 7.4%xliii. Declaye et al.xliv
investigated a modified open drive scroll compressor with R245fa for CHP application and
reported isentropic performance of 75.7% expander, including 2.1 kW shaft power and cycle
performance 8% with condenser temperature is over 50℃, the heat source was hot air and
different heat exchangers were used for the evaporator network.

The micro CHP system was tested in the same way as the hermetic scroll compressor was
converted to act in reverse as an expander and the results showed that the system can generate
6.35% electrical efficiency taking hot water at 65℃, the heat source was a boiler used to heat
up an intermediate heat transfer fluids that were used to boil working fluidsxlv. Miao et al.xlvi
tested organic Rankin cycle system with R123 for 3.25 kW shaft power output with 5.12%
thermal performance at a temperature of 160℃ of heat source. Biomass powered ORC based
micro CHP of 861 W power output, 47.26 kW thermal heat with 1.41% electric power
generation efficiency and 78.7% CHP efficiency reported. The system used biomass boilers
and pressurized water as heat transfer medium at 126℃. HFE7000 was taken as the working
fluid and for expander, a vane type air motor was used. Farrokhi et al.xlvii reported its ORC-
based CHP system, which used R601a (isopentane) and generated 77.4 W of electricity for
water as a heat source at 85℃ temperature with an electrical efficiency of 1.66%.

3.1. ORC technology and applications

The design of the Organic Rankine Cycle is to some degree more straightforward than that of
the steam Rankine cycle: there is no water–steam drum associated with the kettle, and one
single heat exchanger can be utilized to play out the three dissipation stages: preheating,
vaporization and superheating. The varieties of the cycle engineering are likewise more
restricted: warming and turbine draining are by and large not appropriate for the ORC cycle,
however a recuperator can be introduced as fluid preheater between the siphon outlet and the
expander outlet. This permits decreasing the measure of heat expected to disintegrate the liquid
in the evaporator.

The straight forward design can be adjusted and improved relying upon the objective
application. The primary applications are momentarily depicted in the accompanying
segments. Albeit this survey just spotlights on best in class economically accessible ORC
plants, it ought to be noticed that some forthcoming progressed applications for Organic

27 | P a g e
Rankine Cycles are as of now being contemplated, primarily as models or verification of-ideas.
These inventive applications include:

❖ Sun oriented lake power frameworks, in which the ORC framework exploits temperature
inclinations in salt-angle sun-based lakes.

❖ Sunlight based ORC-RO desalination frameworks, where the ORC is utilized to drive the
siphon of an opposite assimilation desalination plant. Sea nuclear power change
frameworks, using the temperature angles (of somewhere around 20 C) in seas to drive a
double cycle.

❖ Cold creation, where the shaft force of the ORC framework is utilized to drive the blower
of a refrigeration framework. Note that this design can likewise be utilized to deliver heat
with a COP41 if the ORC is coupled to a heat siphon.

3.1. Biomass combined heat and power

Biomass is generally accessible in various rural or modern cycles like wood industry or
agrarian waste. Among different means, it very well may be changed over into power by
burning to get heat, which is thusly changed over into power through a thermodynamic cycle.
The expense of biomass is altogether lower than that of petroleum products. However, the
venture important to accomplish clean biomass burning is a higher priority than for exemplary
boilers. For little decentralized units, the age cost of power isn't serious and consolidated heat
and force age is needed to guarantee the productivity of the speculation. Subsequently, to
accomplish high energy change proficiency, biomass CHP plants are generally determined by
the heat request as opposed to by the power request.

The likelihood to utilize heat as a result is a significant resource of biomass ORCs, featuring
the significance of a nearby heat interest, which can be satisfied for example by mechanical
cycles, (for example, wood drying) or space warming (for the most part area warming). Since
heat is somewhat hard to ship across significant distances, biomass CHP plants are more often
than not restricted to 6–10 MW nuclear energy, comparing to 1–2 MW electrical force. This
avoids customary steam cycles that are not practical in this force range.

Improved on graphs of such cogeneration frameworks: heat from the ignition is moved from
the vent gases to the heat move liquid (warm oil) in two heat exchangers, at a temperature
shifting somewhere in the range of 150 and 320 C. The heat move liquid is then coordinated to

28 | P a g e
the ORC circle to dissipate the functioning liquid, at a temperature marginally lower than 300C.
Then, the vanished liquid is extended, goes through a recuperator to preheat the fluid and is at
last consolidated at a temperature around 90 C. The condenser is utilized for boiling water age.

For the specific illustration of Fi, albeit the electrical effectiveness of the CHP framework is
restricted (18%), the general productivity of the framework is 88%, which is a lot higher than
that of brought together force plants, in which the majority of the leftover heat is lost.

To diminish heat misfortunes in the vent gases, these gases should be chilled off to the most
reduced conceivable temperature, to the extent that the corrosive dew point isn't reached. To
accomplish this, two heat move circles are utilized: a high temperature circle and a low
temperature circle. The low temperature circle is introduced after the high temperature circle
on the pipe gases to diminish their outlet temperature.

The primary contending innovation for power age from strong biofuels is biomass gasification:
in this innovation, biomass is changed into an engineered gas made basically out of H2, CO,
CO2 and CH4. This engineered gas is dealt with and sifted to take out strong particles, and is at
long last scorched in an interior burning motor or in a gas turbine.

When looking at the innovation and the expenses of biomass CHP utilizing an ORC with
gasification, it tends to be shown that gasification includes higher speculation costs (about
75%) and higher activity and support costs (about 200%). Then again, gasification yields a
higher capacity to-warm proportion, which makes its abuse more beneficial. It ought to
likewise be noticed that ORC is an all-around demonstrated innovation, while gasification
plants in real activity are for the most part models for exhibit purposes.

Geothermal energy

Geothermal heat sources are accessible over a wide scope of temperatures, from a several
degrees up to 300 C. The real mechanical lower destined for power age is around 80 C: beneath
this temperature the change effectiveness turns out to be too little and geothermal plants are
not prudent. The potential for geothermal energy in Europe and shows that this potential is
extremely high for low temperature sources.

To recuperate heat at a satisfactory temperature, boreholes should by and large be penetrated


in the ground, for the creation well and for the infusion well. The hot brackish water is siphoned
from the previous and infused into the last at a lower temperature. Contingent upon the

29 | P a g e
topographical arrangement, boreholes can be a few thousand meters down, requiring a while
of persistent work. As per Kranz, this prompts a high portion of penetrating expense in the
complete speculation cost (up to 70%) of a geothermal ORC plant. Lazzaretto et al. reports a
substantially more moderate portion of 15.6% for an Italian geothermal double cycle.

Low-temperature geothermal ORC plants are additionally described by generally high helper
utilization: the siphons burn-through from 30% up to over half of the gross yield power. The
principal buyer is the saline solution siphon that needs to course the salt water over huge
distances and with an altogether high stream rate. The functioning liquid siphon utilization is
additionally higher than in higher temperature cycles, in light of the fact that the proportion
between siphon utilization and turbine yield power (''back work proportion'') increments with
diminishing vanishing temperature.

Higher temperature (4150 C) geothermal heat sources empower joined heat and force age: the
consolidating temperature is set to a more elevated level (for example 60 C), permitting the
cooling water to be utilized for region warming. For this situation, the general energy
recuperation productivity is expanded, yet to the detriment of a lower electrical effectiveness.

Solar power plants

Concentrating sun-oriented force is an all-around demonstrated innovation: the sun is followed


and its radiation reflected onto a straight or dependable authority, moving heat to a liquid at
high temperature. This heat is then utilized in a force cycle to create power. The three primary
concentrating sun-based force innovations are the allegorical dish, the sun-oriented pinnacle,
and the illustrative box. Illustrative dishes and sun powered pinnacles are prompt focus
advances, prompting a higher fixation factor and to higher temperatures. The most fitting force
cycles for these advancements are the stirling motor (for limited scope plants), the steam cycle,
or even the consolidated cycle (for sun-based pinnacles).

Allegorical box work at a lower temperature (300–400 C) than point-centered CSP


frameworks. Up to now, they were for the most part coupled to customary steam Rankine
cycles for power age. They are dependent upon similar limits as in geothermal or biomass
power plants: steam cycles require high temperatures, high pressing factors, and hence bigger
introduced ability to be productive.

Natural Rankine Cycles are a promising innovation to diminish venture costs at limited scope:
they can work at lower temperatures, and the absolute introduced force can be downsized to

30 | P a g e
the kW levels. Advances, for example, Fresnel direct concentrators are especially appropriate
for sun-based ORCs since they require a lower speculation cost, yet work at lower temperature.

Up to now, not many CSP plants utilizing ORC are accessible available:

A 1 MWe concentrating sunlight-based force ORC plant was finished in 2006 in Arizona. The
ORC module utilizes n-pentane as the functioning liquid and shows an effectiveness of 20%.
The generally sun oriented to power effectiveness is 12.1% at the plan point. A 100 kWe plant
was dispatched in 2009 in Hawaii by Electro thermal. The heat move liquid temperature in the
authorities is around 120 C.

Some extremely limited scale frameworks are being read for far off-lattice applications, for
example, the verification of-idea kWe framework created for country jolt in Lesotho by ''STG
International''.

Waste heat recovery

Numerous applications in the assembling business reject heat at generally low temperature. In
huge scope plants, this heat is typically overabundant and regularly can't be reintegrated
altogether on location or utilized for area warming. It is hence dismissed to the environment.

This causes two kinds of contamination: poisons (CO2, NOx, SOx, HC) present in the vent
gases produce wellbeing and ecological issues; heat dismissal irritates sea-going harmonies
and negatively affects biodiversity.

Recuperating waste heat mitigates these two kinds of contamination. It can in addition create
power to be burned-through on location or took care of back to the lattice. In such a framework,
the waste heat is generally recuperated by a middle heat move circle and used to dissipate the
functioning liquid of the ORC cycle. A capability of 750 MWe is assessed for power age from
modern waste heat in the US, 500 MWe in Germany and 3000 MWe in Europe (EU-12).

A few ventures present an especially high potential for squander heat recuperation. One model
is the concrete business, where 40% of the accessible heat is ousted through pipe gases. These
vent gases are situated after the limestone preheater or in the clinker cooler, with temperatures
fluctuating between 215 C and 315 C.

CO2 outflows from the concrete business sum for 5% of the all-out world GHG emanations,
and half of it is because of the ignition of non-renewable energy sources in the furnaces.

31 | P a g e
Different models incorporate the iron and steel businesses (10% of the GHG outflow in China
for instance), processing plants and synthetic enterprises.

Heat recovery on internal combustion engines:

An Internal Combustion Engine (ICE) just proselytes around 33% of the fuel energy into
mechanical force on run of the mill driving cycles: an ordinary 1.4 l Spark Ignition ICE, with
a warm productivity going from 15% to 32%, discharges 1.7–45 kW of heat through the
radiator (at a temperature near 80–100 1C) and 4.6–120 kW by means of the fumes gas (400–
900 C).

The heat recuperation Rankine cycle framework (both natural and steam based) is an effective
method for recuperating heat (in correlation with different advancements, for example, thermo-
power and retention cycle cooling). The idea of applying a Rankine cycle to an ICE isn't new
and the main specialized improvements showed up after the 1970 energy emergency. For
example, Mack Trucks planned and constructed a model of such a framework working on the
fumes gas of a 288 HP truck motor. A 450 km on-street test exhibited the specialized
plausibility of the framework and its conservative premium: a decrease of 12.5% in the fuel
utilization was accounted for. Frameworks grew today vary from those of 1970 in light of
advances in the improvement of development gadgets and the more extensive decision of
working liquids. Be that as it may, as of now, no business Rankine cycle arrangement is
accessible.

A large portion of the frameworks being worked on recuperate heat from the exhaust gases and
from the cooling circuit. Paradoxically, the framework created by just recuperates heat from
the cooling circuit. An extra potential heat source is the fumes gas distribution (EGR) and
charge air coolers, in which non-negligible measures of waste heat are disseminated.

The expander yield can be mechanical or electrical. With a mechanical framework, the
expander shaft is straightforwardly associated with the motor drive belt, with a clasp to stay
away from power misfortunes when the ORC influence yield is excessively low. The
fundamental disadvantage of this design is the forced expander speed: this speed is a fixed
proportion of the motor speed and isn't really the ideal speed for amplifying cycle proficiency.
On account of power age, the expander is coupled to an alternator, used to top off the batteries
or supply helper utilities, for example, the cooling. It ought to be noticed that current vehicle

32 | P a g e
alternators show a very low proficiency (around 50–60%), which diminishes the ORC yield
power.

Concerning the expander, the siphon can be straightforwardly associated with the drive belt, to
the expander shaft, or to an electrical engine. Something else, fan utilization can pointedly
lessen the net force yield of the framework.

The presentation of as of late created models of Rankine cycles is promising: the framework
planned by Honda showed a most extreme cycle warm proficiency of 13%. At 100 km/h, this
yields a cycle yield of 2.5 kW (for a motor yield of 19.2 kW) and addresses an increment of
the motor warm proficiency from 28.9% to 32.7%.

A contending innovation under innovative work is the thermoelectric generator (TEG), which
depends on the Seebeck impact: its principle benefits are a generously lower weight than the
ORC framework, and the shortfall of moving parts. Significant downsides are the expense of
materials (which incorporate uncommon earth metals) and the low accomplished proficiency.

ORC manufacturer and market evolution

ORC makers have been available since the start of the 1980s. They give ORC arrangements in
a wide scope of force and temperature levels, as displayed in Table 2. Note that lone producers
with a few business references have been kept in this overview.

The three fundamental producers as far as introduced units and introduced power are Turboden
(Pratt and Whitney) (45% of introduced units around the world, 8.6% of cumulated power),
ORMAT (24% of introduced units, 86% of cumulated force) and Maxxtec (23% of introduced
units, 3.4% of cumulated power). The enormous portion of ORMAT in the cumulated power
is clarified by its emphasis for huge scope, low temperature geothermal twofold plants.

It ought to be noticed that, notwithstanding the producers, numerous organizations are entering
the ORC market with low-limit units, for example for miniature CHP or for WHR on IC motor
exhaust gases. In any case, these organizations have not yet arrived at adequate specialized
development for huge scope serious commercialization.

The last case, the functioning liquid stream rate can be controlled freely, which makes the
guideline of such a framework a lot simpler.

The control of the framework is especially intricate due to the (regularly) transient system of
the warmth source. Notwithstanding, advancing the control is urgent to work on the exhibition

33 | P a g e
of the framework. It is for the most part important to control both the siphon speed and the
expander speed to keep up with the necessary conditions (temperature, pressure) at the
expander gulf.

Another specialized imperative is the warmth dismissal limit. The size of the front warmth
exchanger (either an air-cooled condenser or the radiator associated with a water-cooled
condenser) is restricted by the accessible space and relies upon the presence of a motor radiator,
and perhaps a charge air cooler, an EGR cooler or an air conditioning condenser. The
framework ought to be controlled to such an extent that the dismissed warmth stays inside the
cooling edge, characterized as the cooling limit

The ORC market is developing quickly. Since the primary introduced business ORC plants
during the 1970s, a nearly dramatic development has been expressed, where the advancement
of introduced power and the quantity of plants in activity, in light of an accumulation of
producer information that the ORC is a developed innovation for squander heat recuperation,
biomass CHP and geothermal force, yet is still extremely exceptional for sunlight based
applications. Additionally, frameworks are chiefly introduced in the MW power range and not
many ORC plants exist in the kW power range.

The assortment of ORC modules is enormous and can be ordered by unit size, sort of
innovation, and target application. Some normal ORC module costs, for various applications,
are plotted as an element of their size. Note that the costs are demonstrative just and
incompletely gathered from a non-comprehensive arrangement of ORC makers and from
logical distributions. The dispersing in the information is because of various costs for various
makers, distinctive market methodologies, diverse coordination costs, and so forth.
Accordingly, singular expenses ought not to be summed up, yet are offered simply to delineate
the overall pattern of framework costs comparative with the yield power. For a given objective
application, the expense will in general diminish when the yield power increments. Most
reduced expenses are accounted for squander heat recuperation applications, while geothermal
and CHP plants show higher absolute expense. Complete expense varies from module cost in
that it incorporates designing, structures, and heater (if there should arise an occurrence of
CHP), measure combination, and so on, and can add up to a few times the module cost. These
overflow expenses ought to accordingly never be dismissed while assessing the financial
aspects of an ORC plant.

34 | P a g e
Comparison with the steam Rankine cycle

This part gives a synopsis of the benefits and disadvantages of the ORC innovation. The
intrigued per user can allude to past distributions by a portion of the creators for a more nitty
gritty investigation.

Two fundamental contrasts can be expressed:

1. The incline of the soaked fume bend (right bend of the arch) is negative for water, while
the bend is a lot nearer to vertical for natural liquids. As a result, the impediment of the
fume quality toward the finish of the extension interaction vanishes in an ORC cycle,
and there is no compelling reason to superheat the fume before the turbine channel.
2. The entropy contrast between soaked fluid and immersed fume is a lot more modest for
natural liquids. Thus, the vaporization enthalpy is more modest. Thusly, to take up
equivalent nuclear energy in the evaporator, the natural working liquid mass stream rate
should be a lot higher than for water, prompting higher siphon utilization.

Superheating. As recently expressed, natural liquids ordinarily remain superheated toward the
finish of the development. Hence, there is no requirement for superheating in ORC cycles, in
opposition to steam cycles. The shortfall of buildup additionally decreases the danger of
erosion on the turbine cutting edges, and stretches out its lifetime to 30 years rather than 15–
20 years for steam turbines. Low temperature heat recuperation. Because of the lower limit of
an appropriately chosen natural working liquid, warmth can be recuperated at much lower
temperature (for example with geothermal sources).

Components size. In a steam cycle, the liquid thickness is amazingly low in the low-pressure
part of the cycle. Since pressure drops increment with the square of the liquid speed, the high
volume stream rate requires an increment in the water driven measurement of the channeling
and the size of the warmth exchangers. Likewise, the turbine size is generally corresponding
to the volume stream rate.

Boiler design. ORC cycles empower the utilization of once-box boilers, which dodges steam
drums and distribution. This is because of the somewhat more modest thickness distinction
among fume and fluid for high atomic weight natural liquids. Conversely, the low fume
thickness in steam boilers can create totally different warmth move and pressing factor drop
attributes between fluid water and steam. Complete steam vanishing in a solitary cylinder
should accordingly be stayed away from.

35 | P a g e
Turbine Inlet temperature. In steam Rankine cycles, because of the superheating imperative,
a temperature higher than 450 C is needed at the turbine channel to stay away from drops
development during the extension. This prompts higher warm anxieties in the evaporator and
on the turbine cutting edges and to a greater expense. Siphon utilization. Siphon utilization is
relative to the fluid volume stream rate and to the pressing factor contrast between outlet and
inlet. It tends to be communicated as far as the Back Work Ratio (BWR), which is characterized
as the siphon utilization separated by the turbine yield power. In a steam Rankine cycle, the
water stream rate is somewhat low and the BWR is regularly 0.4%. For a high temperature
ORC utilizing toluene, the run of the mill esteem is 2–3%. For a low temperature ORC utilizing
HFC-134a, values higher than 10% are commonplace. As a rule, the lower the basic
temperature, the higher the BWR.

High pressure factor. In a steam cycle, pressing factors of around 60–70 bar and warm
burdens increment the intricacy and the expense of the steam heater. In an ORC, pressure by
and large doesn't surpass 30 bar. Besides, the functioning liquid isn't vanished straight by the
warmth source (for example a biomass burner) however through a go-between heat move
circle. This makes the warmth recuperation simpler since warm oil can be at surrounding
pressure, and the necessity of an on location steam evaporator administrator is stayed away
from. Gathering pressure. To stay away from air invasion in the cycle, a consolidating pressure
higher than barometrical pressing factor is prudent. Water, be that as it may, has a gathering
pressure for the most part lower than 100 mbar outright. Low temperature natural liquids, for
example, HFC-245fa, HCFC-123 or HFC-134a do meet this necessity. Natural liquids with a
higher basic temperature then again, like hexane or toluene, are sub-atmospheric at
encompassing temperature.

Fluid characteristics. Water as a functioning liquid is extremely helpful contrasted with


natural liquids. Its principle resources are minimal expense and high accessibility, non-
harmfulness, non-combustibility, harmless to the ecosystem (low Global Warming Potential
and invalid Ozone Depleting Potential), compound dependability (no functioning liquid decay
if there should arise an occurrence of problem area in the evaporator), and low consistency
(also, consequently lower grinding misfortunes and higher warmth trade coefficients).
Nonetheless, steam cycles are overall not completely close: water is lost because of breaks,
waste or kettle blow-down. Consequently, a water-treatment framework should be incorporated
with the force plant to take care of the cycle with high-virtue deionized water. A de-aerator

36 | P a g e
should likewise be incorporated to stay away from consumption of metallic parts due the
presence of oxygen in the cycle.

Turbine design. In steam cycles, the pressing factor proportion and the enthalpy drop over the
turbine are both exceptionally high. As a result, turbines with a few development stages are
ordinarily utilized. In ORC cycles, the enthalpy drop is a lot of lower, and single or two-stage
turbines are normally utilized, involving lower cost. Extra results of the lower enthalpy drop
of natural liquids incorporate lower turning velocities and lower tip speed. A lower turning
speed permits direct drive of the electric generator without decrease gear (this is particularly
worthwhile for low force range plants), while the low tip speed diminishes the weight on the
turbine sharp edges and works on their plan. Effectiveness. The proficiency of current high
temperature Organic Rankine Cycles doesn't surpass 24%. Normal steam Rankine cycles show
a warm effectiveness higher than 30%, yet with a more intricate cycle plan (as far as number
of parts or size).

In outline, the ORC cycle is more intriguing in the low to medium force range (commonly not
exactly a couple of MWe), since limited scope power plants can't manage the cost of an on
location administrator, and in light of the fact that it requires basic and simple to fabricate
segments and plan. It is therefore more adjusted to decentralized force age. For high force runs,
the steam cycle is for the most part liked, aside from low temperature heat sources.

Working fluid selection:

The determination of working liquids has been treated in countless logical distributions. Much
of the time, these investigations present a correlation between a bunch of applicant working
liquids as far as thermodynamic execution and in view of a thermodynamic model of the cycle.

While choosing the most suitable working liquid, the accompanying rules and pointers ought
to be considered:

1. Thermodynamic execution: the effectiveness or potentially yield force ought to be just


about as high as feasible for given warmth source and warmth sink temperatures. This
exhibition relies upon various related thermodynamic properties of the functioning
liquid: basic point, acentric factor, explicit warmth, thickness, and so on it's difficult to
build up an ideal for every particular thermodynamic property freely. The most widely
recognized methodology comprises in reenacting the cycle with a thermodynamic
model while benchmarking diverse up-and-comer working liquids.

37 | P a g e
2. Positive or isentropic immersion fume bend: as recently itemized on account of water,
a negative immersion fume bend (''wet'' liquid) prompts beads in the later phases of the
development. The fume should thusly be superheated at the turbine delta to stay away
from turbine harm. On account of a positive immersion fume bend (''dry'' liquid), a
recuperator can be utilized to expand cycle effectiveness.
3. High fume thickness: this boundary is of key significance, particularly for liquids
showing an exceptionally low consolidating pressure (for example silicon oils). A low
thickness prompts a higher volume stream rate: the spans of the warmth exchangers
should be expanded to restrict the pressing factor drops. This unimportantly affects the
expense of the framework. It ought to in any case be noticed that bigger volume stream
rates can permit a more straightforward plan on account of turbo expanders, for which
size is anything but a pivotal boundary.
4. Low consistency: low thickness in both the fluid and fume stages brings about high
warmth move coefficients and low grating misfortunes in the warmth exchangers.
5. High conductivity is identified with a high warmth move coefficient in the warm
exchangers.
6. Acceptable vanishing pressure: as talked about for the instance of water as working
liquid, higher pressing factors typically lead to higher speculation costs and expanded
intricacy.
7. Positive consolidating measure pressure: the low pressing factor ought to be higher than
the environmental pressing factor to stay away from air penetration into the cycle.
8. High temperature solidness: in contrast to water, natural liquids typically experience
compound disintegration and deterioration at high temperatures. The greatest warmth
source temperature is along these lines restricted by the substance steadiness of the
functioning liquid.
9. The liquefying point ought to be lower than the most reduced surrounding temperature
during that time to abstain from freezing of the functioning liquid.
10. High wellbeing level: security includes two principle boundaries—poisonousness and
combustibility. The ASHRAE Standard 34 characterizes refrigerants in wellbeing
gatherings and can be utilized for the assessment of a specific working liquid.
11. Low Ozone Depleting Potential (ODP): the ozone exhausting potential is 11,
communicated as far as the ODP of the R11, set to solidarity.

38 | P a g e
12. Current refrigerants is either invalid or extremely near nothing, since non-invalid ODP
liquids are continuously being eliminated under the Montreal Protocol.
13. Low Greenhouse Warming Potential (GWP): GWP is estimated concerning the GWP
of CO2, picked as solidarity. Albeit a few refrigerants can arrive at a GWP esteem as
high as 1000, there is presently no immediate enactment confining the utilization of
high GWP liquids.
14. Good accessibility and minimal expense: liquids previously utilized in refrigeration or
in the synthetic business are simpler to acquire and more affordable.

39 | P a g e
CHAPTER 4- DATA COLLECTION & DESIGN

The design method considers the heat source and the heat sink temperature. Wang et al.
suggested that thermodynamic parameters such as turbine inlet pressure, turbine inlet
temperature, pinch point temperature difference and condenser temperature have a significant
effect on the external performance of the system. Imran et al. told that evaporation pressure has
a promising effect on thermal efficiency and specific investment costs. For large-scale systems,
for each heat source and the combination of working fluids, there is a maximum vapor
temperature, which is selected as the operating parameter and the expansion machine is
designed or selected accordingly.

In our case, the size of the system is small and it is more economical to get a shelf manufactured
extension machine from the market for our desired power rating. The expansion machine was
selected and the manufacturer set the recommended internal pressure as the vapor pressure.
The expansion ratio was calculated according to the values of the applied condenser and the
vapor pressure. The expansion ratio was calculated according to the values of the applied
condenser and the vapor pressure. The values of the enthalpy difference were calculated for the
expansion ratio and the theoretical power generation capacity was estimated throughout the
expander. Using the anthropic efficiency chart provided by the manufacturer for the expander,
the mass flow rate was calculated to produce 1.11 kW.

The choice of working fluid is a crucial criterion for the performance and reliable operation of
the organic Rankin cycle system. The required properties in ORC working fluid are suitable
thermodynamic properties, high temperature stability, non-toxic, non-corrosion, suitable
pressure range (evaporator and condenser) and environmental protection (ODP, GWP).
Detailed work has been done to investigate the selection of working fluids for the ORC system.
For current work, the selection of working fluids is referred to in published literature, where
thermodynamic simulations were performed with different working fluids to estimate output
power and performance. Environmental hazards and safety have also been a major determining
factor. For the current work, the source side temperature, environmental hazards and
thermodynamic performance have been considered for the selection of the working fluid and
the results of the published literature have been used to review the working fluid options. If the
decision criteria are pure power generation, thermal efficiency, reasonable pressure level and
specific investment costs, R114, R245fa, R123, R601a, R141b and R113 are suitable for sub-
main working fluids. Considering the level of safety and environmental impact, 245fa and

40 | P a g e
R245ca are best suited for waste heat recovery systems. Compared to R245fa, R123, R601a,
R141b and R113, R245fa is non-flammable, it has low toxic levels, has 0 ozone depletion
potential, and is suitable for a wide range of heat temperatures. The R245fa has also been found
to be suitable for the heat source temperature range in the vicinity of 124 C. The available heat
source for this work is also steam in the range of 100-140 C. Thus, R245fa is selected as the
appropriate working fluid for evaporation temperature range, better performance,
environmental and safety considerations.

Following are equations based on thermodynamic analysis that are used to obtain results for
the analysis of data. Heat input rate to the evaporator was calculated using Eq (ii)

𝑄𝑖𝑛 = 𝑚. (ℎ3 − ℎ2 )……………. (ii)

ORC system net power output is calculated by Eq. (iii)

Net Electrical Power (W) = Generator Electrical Power Output-Pump Electrical Power
Consumption …………………….. (iii)

Thermal efficiency of ORC system is given by (iv):

𝐺𝑒𝑛𝑒𝑟𝑎𝑡𝑜𝑟 𝑃𝑜𝑤𝑒𝑟 𝑂𝑢𝑡𝑝𝑢𝑡 (𝑘𝑊)


Efficiency η𝑡ℎ = ………………(iv)
𝑄𝑖𝑛 (𝑘𝑊)

Net efficiency of the ORC system is given by Eq. (v)

𝑁𝑒𝑡 𝑃𝑜𝑤𝑒𝑟 (𝑘𝑊)


Efficiency net η𝑛𝑒𝑡 = ……………………….. (v)
𝑄𝑖𝑛 (𝑘𝑊)

Equipment Details:

The expander machine is a scroll type volumetric expander supplied by Air Square
Manufacturing, Inc. The machine features a 12cc / rev volumetric displacement with a built-in
expansion ratio of 3.5: 1. The machine is capable of providing 1 kW of power output. The
Expander experiment was connected to the AC generator using a magnetic coupling. Together
the AC generators have rated the output at 2.4 kW with a maximum output rating of 3 kW. The
generator provides single phase 120VAC with 60 Hz frequency. The generator has full view
rectifiers and capacitors for self-excitation and is capable of operating without a grid power
supply.

41 | P a g e
The brazed plate type heat exchangers were provided by Janghan Engineers, Inc. as the system
was to be used as a test bed and there is ample scope for use of heat source and sink in different
design conditions. Heat exchangers were able to meet the requirements of the smallest punch
point for the worst performance and for the desired performance. Model BC-50 was provided
as a suitable heat exchanger to operate as a vapor. The heat exchangers have 60 plates with a
total heat transfer area of 6.5 m2. The overall heat transfer coefficient was referred to as 756
W/m2 C. Model BC-50 with 50 number plates was provided as a suitable condenser for our
needs. The heat exchanger covers an area of 5.38 m2. The pressure loss for the working fluid
side is less than 10.6 kPa.

The working fluid pump is an important component of the organic Rankine cycle system. For
small scale systems, the choice of pump is difficult due to the unusual operating conditions. In
the current scenario, the mass flow rate is very small but the differential pressure is very high.
The availability of such pumps in the market is low. General purpose centrifugal pumps are
almost impossible to find in this operating range. Positive displacement pumps are available
options. Diaphragm pumps are known for their ability to produce low flow rates in high
pressure gaps. Pushing in the diaphragm pump is an undesirable phenomenon. A gear pump is
selected as the ultimate solution for their ability to easily deliver fluids at low flow rates in high
pressure gaps. The gear pump has been selected based on practical experience which has helped
in calculating leakage losses through gear clearance. The company recommended a pump
capable of delivering a relatively high flow rate to ensure efficient flow of fluid at the required
flow rate and pressure. This pump was provided by a toothpick pump company. With a
displacement of 2.6 ml / r, the T-Series pump can deliver 3.9 L / min at 3037 rpm while the
required flow rate through the pump can be set to around 1600 rpm. The pump is magnetically
connected to a 0.75 kW 3 phase 380 V motor.

The system is designed for small 1 kW electrical outputs, so instead of using a heavy load bank,
use a single push single throw (SPST) relay from an electric bulb generator of different power
ratings. Were connected 110 V bulbs, which have different power ratings ie 200 W, 100 W, 60
W, 30 W bulbs are used to obtain different phases of resistant load which can be connected to
the generator. If all the bulbs are connected to the generator output circuit, a total load of 1060
W can be connected. The relay is activated by a 24 V signal connected to the control system.

The mass flow through the condenser was made possible using a variable frequency drive
connected to the circulation pump of the cooling tower that was feeding the condenser to the

42 | P a g e
organic Rankin cycle system. The variable frequency drive was also connected to the cooling
tower fan motor which was used to control the temperature of the cold water.

Boilers based on liquefied natural gas (LNG) were used for heating. The boiler had already
been installed in the research facility of the Korea Institute of Energy Research. Maximum
steam generation capacity 500 kg / hr steam flow rate at a pressure range of 1–4 bar. Steam-
using processes fluctuate in steam demand, so excess steam is released into the atmosphere.
With the addition of steam piping networks and control valves, waste steam is sent to the ORC
unit and energy can be used. The current system is a small-scale prototype for exploring the
possibilities of large systems, where no intermediate working fluid is used and the steam is
condensed directly into the ORC evaporator.

Different flow rate meters were used in the system to measure and record the flow rate of
working fluid, heat source, and heat sink fluid. A Coriolis type mass flow meter provided by
KOMETER was used to measure the liquid flow of the working fluid after the pump. For
Coriolis flow meter model KMS-2000 size 15, a connection with 0.2 ٪ full scale accuracy was
used. Mass flow rate can measure mass flow in the range 0-555.6 g / s. A turbine type flow rate
meter KTR-550 provided by size 50 A KOMETER. The flow rate meter can be used to measure
the flow of the vapor phase of the working fluid after the expansion process. The purpose of
using this flow rate meter is to observe the transient response of the flow during the evaporation
procedure for future operations. A Vortex type flow rate meter was also installed to measure
the flow rate of steam. Due to the very low mass flow rate (<40 kg / hr) for this particular
experiment, the flow rate was not reported with confidence. Therefore, it was excluded from
the scope of this work. An electromagnetic type flow rate meter was used to measure the mass
flow rate of cold water in the range of 0-10 m3 / h. This flow rate meter was also provided by
KOMETER company under the specifications of model KTM-800 series which has size 25 A
which can record data with full scale accuracy up to ± 0.5%.

T-type thermocouples with tolerance ± 1 C were used to measure the temperature. The
temperature of the evaporator inlet and exit at working fluid side and expander inlet and outlet
temperature was measured with accuracy class A from Omega via PT100 probe which will
withstand tolerance 0.39 C in the measured temperature range.

Pressure transducers for the 0-16 bar range were supplied by the Huba Control Model 510
Series with 0.5% full-scale tolerance and the transducers for the 0-10 bar range were adopted
from Valcom VPRQ (F) which Can provide 0.8% full scale tolerance.

43 | P a g e
Watt transducers were provided by Light Star with an accuracy class of ± 0.25 of. A 3 phase
transducer was used to measure the power used by the pump motor with a maximum tolerance
of 29 3.29 W.

An optical multimeter tachometer probe was provided by a compact instrument (model A2108
series) to measure the RPM of the generator shaft with a full scale tolerance of 0-6000 RPM ±
0.5%. The probe had an analog output option that was recorded with the data acquisition
system.

Data acquisition and control was done by National Instrumentation cRIO-9074 with 8-slot
integrated 400 MHz real time controller. National instruments C-series input output modules
were used to measure thermocouples, RTDs, pressure transducers, power readings, flow rate
measurements, control valve position and generator speed. The results were electric load relays,
pump variable frequency drives and valve control systems. Lab View Original Software was
used to implement control logic and data retrieval for the test bed.

After setting up the experimental system, the system was run to achieve stable state conditions.
The speed of the working fluid pump was adjusted in such a way that it conveyed the required
mass flow of working fluid through the steamer and bypass pipe. The expander inlet and outlet
valves were initially closed, the steam control valve was slowly opened until the fluid working
on the evaporator's exhaust was flowing in a very hot state. The expander inlet valve was
opened and the bypass valve was closed so that 100% working fluid was flowing through the
expander generator which was connected to the load bank. Stable state conditions were
maintained for different values of power generation. For example, four 800 W bulbs and a 100
W bulb were connected to the generator circuit, so that the imposed power load was 900 W,
then the pump speed was adjusted to get the desired pressure ratio. RPM rating to generate
electrical output so that the expander can rotate. Temperature, pressure, mass flow, electrical
power, pump power consumption and other auxiliary data were recorded. REFPROP software
[44] was used to obtain enthalpy values at the specified temperature and pressure readings.
Table 2 presents the data obtained from the generator for 1000 W, 900 W and 566 W power
generation. For 1000 W generator output, 4 bulbs at 800 W power and 2 bulbs at every 100 W
were connected to the generator circuit. The pump frequency was adjusted and the generator
output as well as the expander speed and working fluid superheating were observed until the
output power from the generator was equal to 1000 W, 900 W or 566 W. To read 566 W, the
generator was connected to 3 bulbs of 200 W. The speed of the pump was different to get the

44 | P a g e
mentioned power output in stable condition. Detailed data of the cycle is presented in Table 2
to provide insights into the cycle behavior at different output power settings. Pure electrical
power was obtained by subtracting the electrical power measured by the pump from the
electrical power generated by the power generator.

45 | P a g e
CHAPTER 5- METHODOLOGY & CALCULATIONS

Fig 6: Proposed ORC system cycle

The proposed ORC system cycle is shown in fig. 1. T-S diagram of this considered ORC system
is represented in fig. 2. The process 1-2 describe work done by pump, 2-3 describe superheating
of working fluid in the evaporator, 3-4 is expansion in a expander, & 4-1 represents removal
of heat in a condenser.

Following are the four processes of Organic Rankine Cycle:

i. Isentropic compression in a pump,


ii. Constant pressure heat addition in a evaporator,
iii. Isentropic expansion in a turbine, &
iv. Constant pressure heat rejection in a condenser

Working fluid enters the pump at state 1 as saturated liquid and is compressed isentropically to
the operating pressure of the evaporator. The working fluid temperature increases somewhat
during this isentropic compression process due to a slight decrease in the specific volume of
working fluid. Working fluid enters the evaporator as a compressed liquid at state 2 and leaves
as a superheated vapor at state 3. The evaporator is basically a large heat exchanger where the
heat originating from combustion gases, nuclear reactors, or other sources is transferred to the
working fluid essentially at constant pressure. The superheated vapor at state 3 enters the
turbine, where it expands isentropically and produces work by rotating the shaft connected to
an electric generator. The pressure and the temperature of fluid drop during this process to the
values at state 4, where fluid enters the condenser. At this state, fluid is usually a saturated

46 | P a g e
liquid–vapor mixture with a high quality. Fluid is condensed at constant pressure in the
condenser, which is basically a large heat exchanger, by rejecting heat to a cooling medium
such as a lake, a river, or the atmosphere. Fluid leaves the condenser as saturated liquid and
enters the pump, completing the cycle. In areas where water is precious, the power plants are
cooled by air instead of water. This method of cooling, which is also used in car engines, is
called dry cooling. Several power plants in the world, including some in the United States, use
dry cooling to conserve water. Remembering that the area under the process curve on a T-S
diagram represents the heat transfer for internally reversible processes, we see that the area
under process curve 2-3 represents the heat transferred to the fluid in the evaporator and the
area under the process curve 4-1 represents the heat rejected in the condenser. The difference
between these two (the area enclosed by the cycle curve) is the Net Work produced during the
cycle.

Fig 7: T-S diagram of ORC system

Following assumptions are taken into consideration while performing calculations:

✓ Heat loss from all the equipment’s and piping are considered negligible.
✓ System is to be considered in ideal cycle i.e the ideal organic Rankine cycle does not
involve any internal irreversibilities.
✓ The system was designed for heat source condition at 2 bar saturated steam.
✓ Mechanical losses in pump and expander and couplings are neglected.

47 | P a g e
✓ Evaporation temperature is taken the maximum attainable temperature of the particular working
fluid.

For R-245fa:

We are using the mass flowrate of 57.2 g/s or 0.572 kg/s.

We take the working fluid in a saturated liquid phase at 100kPa pressure at the pump inlet.

By considering in the ASHARE table of thermodynamic properties of R-245fa:

T1 = 15.14℃,

In the pump, isentropic compression of working fluid take place and pressure increases to 1200
kPa.

The pressure at pump outlet is P2 = 1200 kPa

P2=1200kPa
T1=15.65℃ P1=1200kPa
T1=135℃

P1=100kPa P4=100kPa

T1=15.14℃ T4=39.56℃

Fig 8: ORC cycle using R245fa

Using isentropic relation:

𝛾−1⁄
𝑇2 𝑃2 𝛾
=[ ] ………….. (vi)
𝑇1 𝑃1

By considering in the ASHARE table of thermodynamic properties of R-245fa:

𝛾 = 1.109,

We know

P1 = 100 kPa, &

48 | P a g e
P2 = 1200 kPa.

So, by evaluating (vi) the temperature at pump outlet is obtained as:

T2 = 19.33℃,

In the evaporator heat Q is added in the working fluid by the outer steam and increases its
temperature.

We consider the maximum attainable temperature that R245fa attain in a specific range.

By considering in the ASHARE table of thermodynamic properties of R-245fa:

T3 = 135℃

Now, isentropic expansion takes place in turbine

Using isentropic relation:

𝛾−1⁄
𝑇4 𝑃4 𝛾
=[ ] ………….. (vii)
𝑇3 𝑃3

By considering in the ASHARE table of thermodynamic properties of R-245fa:

𝛾 = 1.840,

We know

P3 = 1200 kPa, &

P4 = 100 kPa.

So by evaluating (vii),

T4 = 39.56℃

By considering in the ASHARE table of thermodynamic properties of R-245fa:

h1 = 219.51 kJ/ (kg.K),

h2 = 225.08 kJ/ (kg.K),

h3 = 488.57 kJ/ (kg.K),

h4 = 433.27 kJ/ (kg.K),

The efficiency of turbine is calculated as:

49 | P a g e
ℎ3 −ℎ4
η𝑡 = …………….(viii)
ℎ3 −ℎ2

η𝑡 = 0.2099

or η𝑡 = 20.99%

The efficiency of overall system is calculated as:

ℎ4 −ℎ1
η𝑡 = 1 − ………………………(ix)
ℎ3 −ℎ2

η𝑠𝑦𝑠 = 0.1988

or η𝑠𝑦𝑠 = 19.88%

For finding rate of heat input, we use following relationship:

𝑄𝑖𝑛 = 𝑚. 𝐶𝑝 (𝑇3 − 𝑇2 )……………… (x)

We know 𝑚. = 0.572 𝑘𝑔/𝑠

By considering in the ASHARE table of thermodynamic properties of R-245fa:

Average value of specific heat between T2 & T3 is 𝐶𝑝 = 1.97.

So by evaluating:

𝑄𝑖𝑛 = 129.98 𝑘𝐽/𝑠

For finding rate of heat input, we use following relationship:

𝑄𝑜𝑢𝑡 = 𝑚. 𝐶𝑝 (𝑇4 − 𝑇1 )…………….. (xi)

Average value of specific heat between T1 & T4 is 𝐶𝑝 = 1.118.

So by evaluating:

𝑄𝑜𝑢𝑡 = 16.06 𝑘𝐽/𝑠

50 | P a g e
For R-123:

We are using the mass flowrate of 57.2 g/s or 0.572 kg/s.

We take the working fluid in a saturated liquid phase at 100kPa pressure at the pump inlet.

By considering in the ASHARE table of thermodynamic properties of R-245fa:

T1 = 27.82℃,

In the pump, isentropic compression of working fluid take place and pressure increases to 1200
kPa.

The pressure at pump outlet is P2 = 1200 kPa

P2=1200kPa
T1=15.65℃ P1=1200kPa
T1=135℃

P1=100kPa P4=100kPa

T1=15.14℃ T4=39.56℃

Fig 9: ORC system using R123

Using isentropic relation:

𝛾−1⁄
𝑇2 𝑃2 𝛾
=[ ]
𝑇1 𝑃1

By considering in the ASHARE table of thermodynamic properties of R-245fa:

𝛾 = 1.108,

We know

P1 = 100 kPa, &

P2 = 1200 kPa.

So, by evaluating the temperature at pump outlet is obtained as:

51 | P a g e
T2 = 35.44℃,

In the evaporator heat Q is added in the working fluid by the outer steam and increases its
temperature.

We consider the maximum attainable temperature that R245fa attain in a specific range.

By considering in the ASHARE table of thermodynamic properties of R-245fa:

T3 = 160℃

Now, isentropic expansion takes place in turbine

Using isentropic relation:

𝛾−1⁄
𝑇4 𝑃4 𝛾
=[ ]
𝑇3 𝑃3

By considering in the ASHARE table of thermodynamic properties of R-245fa:

𝛾 = 1.726,

We know

P3 = 1200 kPa, &

P4 = 100 kPa.

So by evaluating,

T4 = 56.26℃

By considering in the ASHARE table of thermodynamic properties of R-245fa:

h1 = 228.03 kJ/ (kg.K),

h2 = 236 kJ/ (kg.K),

h3 = 463.01 kJ/ (kg.K),

h4 = 415.20 kJ/ (kg.K),

The efficiency of turbine is calculated as:

ℎ3 − ℎ4
η𝑡 =
ℎ3 − ℎ2
52 | P a g e
η𝑡 = 0.2106

or η𝑡 = 21.06%

The efficiency of overall system is calculated as:

ℎ4 − ℎ1
η𝑡 = 1 −
ℎ3 − ℎ2

η𝑠𝑦𝑠 = 0.1755

or η𝑠𝑦𝑠 = 17.55%

For finding rate of heat input, we use following relationship:

𝑄𝑖𝑛 = 𝑚. 𝐶𝑝 (𝑇3 − 𝑇2 )

We know 𝑚. = 0.572 𝑘𝑔/𝑠

By considering in the ASHARE table of thermodynamic properties of R-245fa:

Average value of specific heat between T2 & T3 is 𝐶𝑝 = 1.253.

So by evaluating:

𝑄𝑖𝑛 = 89.27 𝑘𝐽/𝑠

For finding rate of heat input, we use following relationship:

𝑄𝑜𝑢𝑡 = 𝑚. 𝐶𝑝 (𝑇4 − 𝑇1 )

Average value of specific heat between T1 & T4 is 𝐶𝑝 = 1.03.

So by evaluating:

𝑄𝑜𝑢𝑡 = 16.12 𝑘𝐽/𝑠

53 | P a g e
CHAPTER 6 - CONCLUSION
Increasing energy efficiency and reducing environmental impact is an important goal for many
industries around the world, including the cement industry.

We learn about the ORC designing and its processes. In comparison of R123 with R245fa,
R245fa is best for moderate to high temperature waste heat. While, in low temperature waste
heat sources, R123 is a best choice in terms of efficiency.

The total efficiency of an ORC system using R245fa is obtained as 19.88 while efficiency of
turbine is 20.99%. On the other hand, R123 gives the total system efficiency of 17.6% while
efficiency of turbine is 21.06%.

54 | P a g e
REFERENCES

i
UK Government: Climate Change Act. (Accessed 2008) Available at: www.parliament.uk
ii
European Commission. The Paris Protocol—A Blueprint for Tackling Global Climate
Change Beyond 2020; COM (2015)81 Final/2; Communication from the Commission to the
European Parliament and the Council: Brussels, Belgium, 2015.
iii
Alanne, K.; Saari, A. Distributed energy generation and sustainable development. Renew.
Sustain. Energy Rev. 2006, 10, 539–558.
iv
Spigarelli, F.; Curran, L.; Arteconi, A. China and Europe’s Partnership for a More
Sustainable World: Challenges and Opportunities; Emerald Group Publishing Limited:
Bingley, UK, 2016.
v
Larjola, J.: Electricity from Industrial Waste Heat using High-Speed Organic Rankine Cycle
(ORC). International Journal of Production Economics 41, 227-235 (1995).
vi
Tartiere, T., Astolfi, m.: A World Overview of the Organic Rankine Cycle Market. Energy
Procedia 129, 2-9 (September 2017)
vii
Energy, E.: The Potential for Recovering and Using Surplus Heat from Industry., UK
Department of Energy and Climate Change, London.
viii
Cayer, E., Galanis, N., Nesreddine, H.: Parametric Study and Optimisation of a Transcritical
Power Cycle Using a Low-Temperature Heat Source. Applied Energy 87, 1349-1367.
ix
Hung, T.-C.: Waste Heat Recovery of Organic Rankice Cycle Using Dry Fluids. Energy
Conversion and Management 42(5), 539-553
x
Wang T, Zhang Y, Peng Z, Shu G. A review of researches on thermal exhaust heat recovery
with Rankine cycle. Renew Sustain Energy Rev 2011; 15: 2862–71.
http://dx.doi.org/10.1016/j.rser.2011.03.015
xi
Tchanche BF, Lambrinos Gr, Frangoudakis A, Papadakis G. Low-grade heat conversion into
power using organic Rankine cycles – A review of various applications. Renew Sust Energ
Rev 2011; 15(8):3963–79. http://dx.doi.org/10.1016/j.rser.2011.07.024
xii
Hung TC, Shai TY, Wang SK. A review of organic Rankine cycles (ORCs) for the recovery
of low-grade waste heat. Energy 1997; 22: 661–7. http://dx.doi.org/10.1016/S0360-
5442(96)00165-X
xiii
Lecompte S, Huisseune H, van den Broek M, Vanslambrouck B, De Paepe M. Review of
organic Rankine cycle (ORC) architectures for waste heat recovery. Renew Sustain Energy
Rev 2015; 47:448–61. http://dx.doi.org/10.1016/j.rser.2015.03.089
xiv
Bundela. Sustainable development through waste heat recovery. Am J Environ Sci 2010;
6:83–9 http://dx.doi.org/10.3844/ajessp.2010.83.8
xv
Imran M, Park B-S, Kim H-J, Lee D-H, Usman M. Economic assessment of greenhouse gas
reduction through low-grade waste heat recovery using organic Rankine cycle (ORC). J Mech
Sci Technol 2015; 29: 835–43. http://dx.doi.org/10.1007/s12206-015-0147-5
xvi
Campana F, Bianchi M, Branchini L, De Pascale A, Peretto A, Baresi M, et al. ORC waste
heat recovery in European energy intensive industries: energy and GHG savings. Energy
Convers Manage 2013; 76:244–52. http://dx.doi.org/10.1016/j.enconman.2013.07.041
xvii
Cavazzini G, Dal Toso P. Techno-economic feasibility study of the integration of a
commercial small-scale ORC in a real case study. Energy Convers Manage 2015; 99:161–75.
http://dx.doi.org/10.1016/j.enconman.2015.04.04
xviii
S.K. Ray, G. Moss, Fluorochemicals as working fluids for small Rankine cycle power units,
Advanced Energy Conversion 6 (1966) 89–102
xix
S. Devotta, F.A. Holland, Comparison of theoretical Rankine power cycle performance data
for 24 working fluids, Heat Recovery Systems 5 (6) (1985) 503–510.

55 | P a g e
xx
R. DiPippo, Geothermal power plants: Principles, Applications and Case Studies, Elsevier
Advanced Technology, Oxford, 2005.
xxi
V. Maizza, A. Maizza, Unconventional working fluids in organic Rankine-cycles for waste
energy recovery systems, Applied Thermal Engineering 21 (2001) 381–390.
xxii
B.-T. Liu, K.-H. Chien, C.-C. Wang, Effect of working fluids on organic Rankine cycle for
waste heat recovery, Energy 29 (2004) 1207–1217
xxiii
Y.S.H. Najjar, Efficient use of energy by utilizing gas turbine combined systems, Applied
Thermal Engineering 21 (2001) 407– 438.
xxiv
Rayegan R, Tao YX. A procedure to select working fluids for solar organic Rankine cycles
(ORCs). Renew Energy 2011; 36:659–70. http://dx.doi.org/10.1016/j.renene.2010.07.010
xxv
Hung TC, Wang SK, Kuo CH, Pei BS, Tsai KF. A study of organic working fluids on system
efficiency of an ORC using low-grade energy sources. Energy 2010; 35:1403–11.
http://dx.doi.org/10.1016/j.energy.2009.11.025
xxvi
Walraven D, Laenen B, D’haeseleer W. Comparison of thermodynamic cycles for power
production from low-temperature geothermal heat sources. Energy Convers Manage 2013;
66:220–33. http://dx.doi.org/10.1016/j.enconman.2012.10.00
xxvii
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.
http://dx.doi.org/10.1016/j.energy.2004.01.004
xxviii
Li Y-R, Du M-T, Wu C-M, Wu S-Y, Liu C, Xu J-L. Economical evaluation and
optimization of subcritical organic Rankine cycle based on temperature matching analysis.
Energy 2014; 68:238–47. http://dx.doi.org/10.1016/j.energy.2014.02.038
xxix
Xu J, Yu C. Critical temperature criterion for selection of working fluids for subcritical
pressure Organic Rankine cycles. Energy 2014; 74:719–33.
http://dx.doi.org/10.1016/j.energy.2014.07.038
xxx
Wang X, Liu X, Zhang C. Parametric optimization and range analysis of organic Rankine
cycle for binary-cycle geothermal plant. Energy Convers Manage 2014; 80:256–65.
http://dx.doi.org/10.1016/j.enconman.2014.01.026
xxxi
Quoilin S, Declaye S, Tchanche BF, Lemort V. Thermo-economic optimization of waste
heat recovery organic Rankine cycles. Appl Therm Eng 2011; 31:2885–93.
http://dx.doi.org/10.1016/j.applthermaleng.2011.05.014
xxxii
Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low temperature organic
Rankine cycles. Energy 2007; 32:1210–21. http://dx.doi.org/10.1016/j.energy.2006.07.001
xxxiii
Bao J, Zhao L. A review of working fluid and expander selections for organic Rankine
cycle. Renew Sustain Energy Rev 2013; 24:325–42.
http://dx.doi.org/10.1016/j.rser.2013.03.040
xxxiv
Qiu G, Liu H, Riffat S. Expanders for micro-CHP systems with organic Rankine cycle.
Appl Therm Eng 2011; 31:3301–7. http://dx.doi.org/10.1016/j.applthermaleng.2011.06.008
xxxv
Gnutek Z, Kolasin´ ski P. The application of rotary vane expanders in organic rankine cycle
systems—thermodynamic description and experimental results. J Eng Gas Turbines Power
2013; 135:061901. http://dx.doi.org/10.1115/1.4023534
xxxvi
Lemort V, Quoilin S, Cuevas C, Lebrun J. Testing and modeling a scroll expander
integrated into an organic Rankine cycle. Appl Therm Eng 2009; 29:3094–102.
http://dx.doi.org/10.1016/j.applthermaleng.2009.04.01
xxxvii
Lemort V, Declaye S, Quoilin S. Experimental characterization of a hermetic scroll
expander for use in a micro-scale Rankine cycle. Proc Inst Mech Eng Part A J Power Energy
2012; 226:126–36. http://dx.doi.org/10.1177/0957650911413840
xxxviii
Chang J-C, Hung T-C, He Y-L, Zhang W. Experimental study on low temperature organic
Rankine cycle utilizing scroll type expander. Appl Energy 2015; 155:150–9.
http://dx.doi.org/10.1016/j.apenergy.2015.05.118
56 | P a g e
xxxix
Declaye S, Quoilin S, Guillaume L, Lemort V. Experimental study on an open drive scroll
expander integrated into an ORC (organic Rankine cycle) system with R245fa as working fluid.
Energy 2013; 55:173–83. http://dx.doi.org/10.1016/j.energy.2013.04.003
xl
Song P, Wei M, Shi L, Danish SN, Ma C. A review of scroll expanders for organic Rankine
cycle systems. Appl Therm Eng 2014; 75:54–64.
http://dx.doi.org/10.1016/j.applthermaleng.2014.05.094
xli
AIR SQUARED, INC. www.airsquared.com [accessed 05.12.15]
xlii
Pei G, Li J, Li Y, Wang D, Ji J. Construction and dynamic test of a small-scale organic
Rankine cycle. Energy 2011; 36:3215–23. http://dx.doi.org/10.1016/j.energy.2011.03.010.
xliii
Pei G, Li J, Li Y, Wang D, Ji J. Construction and dynamic test of a small-scale organic
Rankine cycle. Energy 2011; 36:3215–23. http://dx.doi.org/10.1016/j.energy.2011.03.010
xliv
Declaye S, Quoilin S, Guillaume L, Lemort V. Experimental study on an open drive scroll
expander integrated into an ORC (organic Rankine cycle) system with R245fa as working fluid.
Energy 2013; 55:173–83. http://dx.doi.org/10.1016/j.energy.2013.04.003
xlv
Oudkerk J-F, Quoilin S, Declaye S, Guillaume L, Winandy E, Lemort V. Evaluation of the
energy performance of an organic Rankine cycle-based micro combined heat and power system
involving a hermetic scroll expander. J Eng Gas Turbines Power 2013; 135:042306.
http://dx.doi.org/10.1115/1.4023116
xlvi
Qiu G, Shao Y, Li J, Liu H, Riffat SB. Experimental investigation of a biomass fired ORC-
based micro-CHP for domestic applications. Fuel 2012; 96:374–82.
http://dx.doi.org/10.1016/j.fuel.2012.01.028
xlvii
Farrokhi M, Noie SH, Akbarzadeh AA. Preliminary experimental investigation of a natural
gas-fired ORC-based micro-CHP system for residential buildings. Appl Therm Eng 2013.
http://dx.doi.org/10.1016/j.applthermaleng.2013.11.060

57 | P a g e

You might also like