Download as pdf or txt
Download as pdf or txt
You are on page 1of 70

LOW TEMPERATURE POWER GENERATION USING HFE-7000 IN A

RANKINE CYCLE

_______________

A Thesis

Presented to the

Faculty of

San Diego State University

_______________

In Partial Fulfillment

of the Requirements for the Degree

Master of Science

in

Mechanical Engineering

_______________

by

Austin D. Reid

Fall 2010
iii

Copyright © 2010

by

Austin D. Reid

All Rights Reserved


iv

ABSTRACT OF THE THESIS

Low Temperature Power Generation Using HFE-7000 in a


Rankine Cycle
by
Austin D. Reid
Master of Science in Mechanical Engineering
San Diego State University, 2010

This thesis presents the modeling, construction, and testing of a cost-effective Rankine
cycle that utilizes a low-grade heat source (100°C); a non-toxic, environmentally benign
working fluid (heptafluoropropyl methyl ether, or HFE-7000); and a scroll-type expander to
generate electric power. A thermal efficiency of 3.1% was achieved, and a clear path
forward has been presented to achieve 9.7% efficiency. This experiment validates the use of
HFE-7000 as a working fluid in a Rankine cycle – both in a custom application or as a drop-
in replacement for less-desirable fluids. Such a system has potential uses that include
bottoming cycles for industrial processes and inexpensive non-concentrating solar-thermal
power plants.
v

TABLE OF CONTENTS

PAGE

ABSTRACT ............................................................................................................................. iv
LIST OF TABLES .................................................................................................................. vii
LIST OF FIGURES ............................................................................................................... viii
LIST OF ABBREVIATIONS .................................................................................................. ix
ACKNOWLEDGEMENTS ..................................................................................................... xi
CHAPTER
1 INTRODUCTION .........................................................................................................1 
Review of Literature ................................................................................................2 
Proposed Research ...................................................................................................7 
2 SYSTEM MODELING .................................................................................................8 
3 SYSTEM DEVELOPMENT .......................................................................................14 
Heat Source ............................................................................................................15 
Steam................................................................................................................15 
Internal Combustion Engine ............................................................................15 
Water Heater Source ........................................................................................16 
Heat Rejection Sink ...............................................................................................18 
System Components...............................................................................................19 
Heat Exchangers ..............................................................................................19 
Pump ................................................................................................................19 
Expander ..........................................................................................................20 
Multi-Vane Expander.......................................................................................20 
Screw Expander ...............................................................................................21 
Scroll Expander ................................................................................................23 
Turbo Expander ...............................................................................................25 
Expander Selection ..........................................................................................26 
Compressor Conversion ...................................................................................27 
Plumbing Configuration.........................................................................................31 
vi

Electrical Load .......................................................................................................33 


Measurement ..........................................................................................................35 
Working Fluid ........................................................................................................36 
Performance .....................................................................................................36 
Environmental Impact ......................................................................................37 
Safety ...............................................................................................................39 
Fluid Selection .................................................................................................39 
4 RESULTS AND DISCUSSION ..................................................................................41 
Experimental Data .................................................................................................41 
Update to Theoretical Model .................................................................................42 
Data Comparison ...................................................................................................43 
Discussion ..............................................................................................................44 
Pump Performance ...........................................................................................44 
Expander ..........................................................................................................46 
Heat Source and Sink .......................................................................................47 
Overall Efficiency Gains..................................................................................47 
HFE-7000 .........................................................................................................48 
Cost Analysis ...................................................................................................49 
5 CONCLUSIONS..........................................................................................................51 
Summary ................................................................................................................51 
Recommendations ..................................................................................................52 
REFERENCES ........................................................................................................................53
APPENDIX
SCROLL DEVICE PROPERTIES ..............................................................................57 
vii

LIST OF TABLES

PAGE

Table 1. Theoretical State Calculations ...................................................................................12 


Table 2. Expander Comparison Summary ...............................................................................26 
Table 3. Environmental Factors for Possible Working Fluids .................................................39 
Table 4. Summary of Experimental Data ................................................................................41 
Table 5. Experimental State Calculations ................................................................................43 
Table 6. Optimized State Calculations.....................................................................................48 
Table 7. Comparison of Refrigerant Density ...........................................................................49 
Table 8. System Cost Breakdown ............................................................................................49 
viii

LIST OF FIGURES

PAGE

Figure 1. System schematic. ......................................................................................................8 


Figure 2. P-H diagram for HFE-7000. .......................................................................................9 
Figure 3. P-H diagram of theoretical ORC states. ...................................................................13 
Figure 4. ORC system at outset of project. ..............................................................................14 
Figure 5. Heat source installed.................................................................................................17 
Figure 6. Heat sink installed. ...................................................................................................18 
Figure 7. Multi-vane expander. ................................................................................................20 
Figure 8. Screw expansion device (oil injected). [28] .............................................................22 
Figure 9. Scroll expander operation. ........................................................................................23 
Figure 10. Expander schematic. ...............................................................................................27 
Figure 11. Expander with top removed and check valve assembly. ........................................28 
Figure 12. Scroll device sides A and B. ...................................................................................28 
Figure 13. Compressor flow diagram. .....................................................................................29 
Figure 14. Expander flow diagram. .........................................................................................30 
Figure 15. Scroll device installed with inlet machined. ...........................................................31 
Figure 16. Cap before and after modifications. .......................................................................31 
Figure 17. Expander installed in ORC system. ........................................................................32 
Figure 18. Pump to pre-heater link before and after modification. .........................................33 
Figure 19. Run capacitor wiring. .............................................................................................34 
Figure 20. Electrical setup diagram. ........................................................................................35 
Figure 21. 3-phase load with clamp-on ammeter and switches. ..............................................35 
Figure 22. P-H diagram of actual ORC states..........................................................................44 
Figure 23. Pump performance curve. .......................................................................................45 
Figure 24. Compressor performance data under various conditions. ......................................58 
Figure 25. Compressor specifications. .....................................................................................59 
ix

LIST OF ABBREVIATIONS

ALT – Atmospheric Life Time


ORC – Organic Rankine Cycle
GWP – Global Warming Potential
ODP – Ozone Depletion Potential
STP – Standard Temperature and Pressure
GPM – Gallons per Minute
CFH – Cubic Feet per Hour
HFC - Hydrofluorocarbon
HFE – Hydrofluoroether
HCFC – Hydrochlorofluorocarbon
LFL –Lower Flammability Limit
PV - Photovoltaic
P, p – Pressure (kilopascals)
T – Temperature (F, C, and K alternately)
H, h – Enthalpy (kJ/kg)
– Heat Transfer (kW)
– Work (kW)
S, s – Entropy (J/mol-K)
η – Efficiency (%)
, mdot – Mass flow rate (kg/s)
cp – Heat capacity (J/kg-K)

Subscripts:
1,2,3,4 – States
e - expander
gen – generator
H2O - Water
x

in – inlet, supply
out – outlet, sink
p – pump
th – thermal
xi

ACKNOWLEDGEMENTS

I wish to acknowledge the work of Terry Ireland, Mike Lester, Bill Lekas, and everyone
at the SDSU Physical Plant that helped make the hardware functional.
1

CHAPTER 1

INTRODUCTION

Many industrial processes produce waste heat that is typically rejected to a lower
temperature heat sink, but can instead be recovered to produce useful energy. This waste
energy is often rejected at lower temperatures than would be useful for typical energy
conversion systems. Recovery of waste heat offers the benefit of increasing overall
efficiency in the case of a power generation, or of providing auxiliary power in other waste
heat application. The standard Rankine Cycle, often used for primary power generation
using a heat source, operates at relatively high temperatures (250-600°C) in order to
maximize Carnot efficiency, rendering many lower-temperature heat sources useless as
energy sources. However Organic Rankine Cycles (ORC), which use organic working fluids
rather than water or other fluids, can operate at low enough temperatures to take advantage of
lower-temperature sources.
In addition to industrial waste heat, solar sources are also wont to produce lower-
temperature energy. To achieve high-temperature from a solar source typically requires high
concentration ratios produced by large-area, and thus expensive, reflective concentrators. A
lower temperature requirement allows a solar collector to be a smaller or less efficient
concentrator, or possible even a direct-heated system without any concentrating at all. The
latter would be an extremely low-cost heat source.
The state of the art in ORCs has several limitations. Many of the working fluids are
ozone depleting (HCFC fluids) or toxic. Components are expensive relative to power
produced – a turbine expander is appropriate for use in a steam Rankine Cycle that produces
electric power in the MW range, but is too costly for a system that only produces power on
the kW scale. Further, the minimum working temperature of many cycles that have been
investigated, often above 200°F (Dai, Wang, & Gao, 2009; Hettiarachchi, Golubovic, Worek,
& Ikegami, 2007; Hung, Shai, & Wang, 1997), is still too high for many applications,
including some waste heat recovery and solar sources.
2

REVIEW OF LITERATURE
Much research has been done to investigate the use of Rankine cycles to capture low-
quality heat. While alternative fluids have been investigated, such as Zeotropic mixtures
(Wang & Zhai, 2009), the vast majority of the research focuses on organic working fluids.
Organic working fluids offer better performance at lower temperatures than typical inorganic
working fluids, namely water (Dai et al., 2009; Yamamoto, Furahata, Arai, & Mori, 2001).
Many of the studies performed are purely theoretical, with no real-world validation of
thermodynamic models. Hettiarachchia et al. (2007) evaluated design criteria such as
working fluid, evaporation and condensation temperatures, and system cost to optimize the
ORC for geothermal sources, concluding that ammonia is the primary candidate, followed by
HCFC 123 for heat sources nearing 100°C. Liu, Chien, and Wang (2004) performed an
analysis of several working fluids at various evaporator temperatures and found that
evaporator efficiency is maximized by fluids with low enthalpies of vaporization, and thus
these fluids are most appropriate for an ORC. Mago, Srinivqasa, Chamra, and Somayaji
(2008) used an exergy analysis to increase theoretical efficiency in an ORC using
regenerative heating, concluding that the evaporator contributes most to exergy destruction in
an ORC.
Some research has been done to improve the accuracy of existing models using
computational and finite element methods. This higher resolution can be used to improve the
design of a Rankine system without incurring the cost of a real-world experiment. Vaja and
Gambarotta (2010a; 2010b) performed extensive modeling of the ORC system, producing
complete Simulink libraries of components for future system modeling. Particular focus was
placed on the evaporator and condenser, resulting in an accurate parametric finite difference
model. Their model consists of “state determined” (evaporator and condenser) and “not state
determined” (pump and expander) to avoid algebraic loops and iterative numerical solutions.
A set of detailed simulation results is produced for HCFC123, but no validation has been
done to date. The theoretical system operates at 200°C, producing greater that 275kW at
18% thermal efficiency.
Toffolo, Lazzarettoa, Manentea, and Rossi (2010) took a similar approach by
decoupling the heat transfer portions of the cycle (evaporator and condenser) from the other
components, and using the “HEATSEP” model to characterize the heat transfer. Working
3

fluids isobutene and R134a were considered with heat source temperatures down to 130°C,
and R134a was found to have higher exergy recovery coefficients.
Much research has gone into the selection of working fluids for various cycles based
on fluid characteristics. One such characteristic is the fluid’s saturated vapor curve. Many
papers focus solely on “dry” working fluids – fluids for which the saturated vapor curve is
positive in a temperature-entropy (T-s) diagram. Dry fluids are preferable to “wet” fluids
because they do not condense after the sudden enthalpy drop through the expander.
Isentropic fluids have an approximately vertical vapor saturation curve. Hung et al. (1997)
analyzed a series of wet, dry, and isentropic fluids over various temperatures and pressures,
and more recently Hung (2001) focused on dry fluids in “Waste heat recovery of organic
Rankine cycle using dry Fluids,” concluding that that isentropic fluids are most appropriate
for use with low temperature heat sources. Mago et al. (2008) also focused on dry fluids,
while still incorporating regenerative heating, in “An examination of regenerative organic
Rankine cycles using dry fluids”, concluding that regeneration is appropriate for an ORC,
increasing efficiency and decreasing irreversibilities. Mago et al. also show that superheating
dry fluids adds irreversibility to the system – with no risk of condensation in a dry fluid, the
fluid should enter the expander as saturated vapor rather than a superheated vapor.
In “Investigation of the criteria for fluid selection in Rankine cycles for waste heat
recovery”, Siddiqi and Atakan (2010) suggest that T−H diagrams are the best and most
intuitive way to judge the performance or the weakness a fluid for a particular heat source
and heat sink. However many of the fluorinated hydrocarbons were found to be
inappropriate for the high temperature heat sources considered in the research.
Many fluids are common for use in ORCs, but some papers investigate novel or
previously overlooked fluids. Much of this is driven by environmental laws that phase out
heritage fluids. In “HFC-245fa Working Fluid in Organic Rankine Cycle - A Safe and
Economic Way to Generate Electricity from Waste Heat”, Zyhowski, Brown, and Achaichia
(2010) evaluate HFC-245fa as a potential safe and benign working fluid. They conclude that
the fluid could perform efficiently in an ORC due to its high heat capacity and gas density.
The fluid is found to outperform water in systems operating with an approximately 150°C
heat source and creating 50kW.
4

Husband and Beyene (2008) address environmental effects and component costs in
“Low-grade heat-driven Rankine cycle, a feasibility study.” The paper demonstrates the
feasibility of a low-grade heat recovery system that can produce electrical power using a
relatively benign working fluid. Specifically, a hydrofluroether – heptafluoropropyl methyl
ether (HFE-7000 or Novec 7000) has low global warming potential (GWP), zero ozone
depletion potential (ODP), low ALT, low toxicity, and zero flammability. A theoretical
efficiency of 11% for a 10kW system was shown. Further, a scroll device was suggested as
an expander.
The type of expander used in the cycle is another common area of research. Papers
have focused on several common expanders, including multi-vane, which performs well and
is low-cost (O’Callaghan, Badr, Probert, Bell, & Patel, 1985); turbo, which is high cost and
high performing, but largely unproven for low temperature power generation (Teagan &
Clay, 1973); and screw, which performs well with wet fluids though the cost remains
somewhat high (Smith, Stosic, & Kovacevic, 1999). The benefits and drawbacks of each of
these devices are described in further detail in the Expander section below.
Some novel devices have been investigated as well. Giampaolo and Stefano (2010)
research the Wankel device in “Model of a steam Wankel Expander.” Though in its infancy
in this application, the Wankel expander shows some promise in low-temperature
applications, achieving a theoretical system efficiencies greater than 10% with water working
fluid at 150°C.
Another device investigated by Giampaolo and Leonardo (2010) is a reciprocating
engine operating with a heat source between 100 and 150°C. Results of the model indicate
that the expansion device may be promising in a low-pressure (7 bar) steam application
producing power up to 10kW.
The scroll device is one particular expander that has been the subject of myriad
research, both theoretical and experimental. Clemente, Micheli, Reini, and Taccani (2010)
have developed a numerical model of a scroll device capable of estimating the performance
of the device as both expander and compressor. Theoretical calculations have been
compared to a set of experimental data which validate the accuracy of the model for a range
of input parameters such as expansion ratio and shaft speed, as well as specific common
working fluids.
5

Yanagisawa, Fukuta, Ogo, and Hikichi (2001) found that the most dominant factor
affecting scroll expander efficiency is not leakage loss but mechanical loss from the orbiting
motion; pressure in the scroll pocket drops due to a throttling effect when the wrap opens
before the expansion occurs.
Zanelli and Favrat (1994) detail the conversion of a 1 to 3.5 kWe scroll expander
from a compressor, and the use of an organic Rankine cycle test facility for expanders up to
10 kWe. Expander efficiencies reached with HFC-134a reach 65% with shaft speeds
between 2400 and 3600 RPM.
Low shaft research has been performed by Kim, Ahn, Park, and Rha (2007),
achieving experimental and theoretical efficiencies of 34 and 65% respectively at shaft
speeds below 1400 RPM.
Kane, Favrat, Gay, and Andres (2007) measured efficiency of an ORC at about 7%
with a heat source at around 90°C for a 7kWe system. This research, though large in scale,
has many of the features in common with research presented in this thesis.
Saitoh, Yamada, and Wakashima (2007) report a total thermal efficiency of 7% in a
solar powered system using a scroll expander and a concentrating parabolic reflector.
Peterson, Wang, and Herron (2008) research a lower power system that produces
from 187 to 256 W. At nominal ambient outside temperatures (22.5 °C), the system
efficiency was 7.2%. The authors found that the expander limited the overall system
efficiency with efficiencies between 45 and 50%.
Lemort and Quoilin of the University of Liege, Belgium are responsible for
significant experimental work, performing extensive testing with ORC systems utilizing a
scroll expansion device. Between two studies (Lemort & Quoilin, 2009; Quoilin, Lemort, &
Lebrun, 2010), over 40 steady state performance points were gathered for HCFC-123 and
water evaluating the effects of boiler temperature, mass flow rate, condenser temperature,
and expander rotational speed on shaft power, the expander isentropic effectiveness, and the
cycle efficiency. Expander efficiencies ranging from 42% to 68% were observed yielding
system efficiency up to 7.4% with heat source temperatures down to 120°C. The research
concludes that HCFC-123 offers superior performance with the scroll device relative to water
due to higher volumetric performance and the lower under-expansion losses. The authors
further recommend the use of HFC-245fa as HCFC-123 is phased out.
6

Overall, it is observed that much research has been performed with scroll expansion
devices, with a wide range of input temperatures, fluids, and system sizes. Experimental
ORC systems of the approximate size and temperature range as the system of interest in this
thesis report efficiencies between 7 and 8%.
Additional validation and real-world experimentation has been done to investigate the
operation of the ORC. Yamamoto et al. (2001) created hardware to validate their fluid of
choice, HCFC-123, versus water. They were able to achieve 1.25% efficiency. An intensive
empirical study was performed by Quoilin (2007) in “Experimental Study and Modeling of a
Low Temperature Rankine Cycle for Small Scale Cogeneration.” Several variables were
considered, including hot air source temperature, expander rotational speed, and refrigerant
charge. Refrigerant charge was found to have a significant impact on the performance of the
cycle.
One commercial system that has been installed is a low-temperature but large-scale
power plant installed at Chena hot springs in Alaska (Cogswell, 2006). Operating with a heat
source of only 73°C, the system uses turbo expanders to generate 200kW at a net thermal
efficiency of over 8%. Further, the cost of this system is minimized by using commercial
refrigeration components. The expansion device is a converted in-line centrifugal
compressor.
Several papers have specifically investigated the cost effectiveness of ORC systems.
Husband and Beyene (2008) perform component research and economic considerations and
conclude that a 10kW solar driven ORC may be cost competitive when weighed against a
comparable PV system.
Tchanche, Quoilin, Declaye, Papadakis, and Lemort (2010) perform an exhaustive
cost and performance analysis for a variety of fluids and parameters in “Economic
Optimization of Small Scale Organic Rankine Cycles.” The authors found a strong
correlation between pressure ratio and both cost and power output, however the respective
minimum and maximum did not match. The operating point for maximum power does not
match that for minimum investment cost. The total investment cost includes costs of a scroll
expander, evaporator, condenser, fluid pump, pipes, working fluid charge, water cooling
pump, liquid reservoir, control system, miscellaneous hardware and a labor cost equal to 10%
7

of the total material cost. Many fluids were considered including HCFC-123 and HFC-
245fa, which both achieve investment costs below 5000 Euro/kWe (6500 $/kWe).
Smith, Stosic, Kovacevic, and Langson (2007) also performed a cost analysis for a
larger system, up to 20kW, and found that the installed cost per kW was as low as
2500$/kWe, much lower than the cost estimated by Tchanche, but at a much larger scale.
Such a system was found to be approximately 30% cheaper than a 200kW turbo expander
system quoted for the research, and could continue to be cheaper up to 500kW. The research
investigated heat source temperatures down to 90°C, and focused on the screw expansion
device.
The primary goal of the Chena project, and one which was achieved, was to reduce
the cost of the installed ORC system to 1300 $/kW (Cogswell, 2006).
All of these areas: fluid selection, expansion device, theoretical modeling, experimental
demonstration, and cost effectiveness will be considered in this thesis.

PROPOSED RESEARCH
This research will build upon the work done by Husband and Beyene. Using the
theoretical model as a starting point, an existing system optimized for use with HCFC-134
working fluid will be modified to run with HFE-7000. The operational system will be
measured and analyzed to establish overall system efficiency, sources of irreversibility, and
potential areas for improvement.
Successful research will provide four immediate benefits:
 The use of an HFE fluid in an ORC will be demonstrated for the first time.
 The use of a low-temperature power source (less than 100°C) will be demonstrated.
 The existing mathematical model for HFE-7000 will be validated or modified,
allowing future systems to be designed using more ideal parameters such as
temperature and pressure, or other components.
 Inexpensive components will be used to demonstrate cost effectiveness of the ORC
relative to photovoltaic systems.
This research will require procurement of components and materials (including HFE-
7000), construction and assembly of the system, leak-checking and fluid-charging of the
system, location and connection of heat source and sink, and finally taking of temperature,
pressure, and electric measurements during system operation.
8

CHAPTER 2

SYSTEM MODELING

The theoretical analysis of the fluid is a state analysis – at certain points in the system
the state of the fluid is calculated. Each state includes all properties of the fluid such as
temperature, pressure, enthalpy, entropy, and specific volume, but can be described with only
two – in this case, temperature and pressure. Through each component, certain properties of
the fluid change. Through the pump, pressure increases; through the condenser, temperature
decreases; etc.
Although the analysis is theoretical, a system has already been assembled, so certain
component properties are already known. This differs from a typical analysis in that
properties are imposed on the model rather than using the model to select components with
ideal properties. The system schematic can be seen in Figure 1, with the states labeled.

Figure 1. System schematic.


9

HFE-7000 is unique in the ORC application because of its high boiling point. A
typical refrigerant used in a similar application, such as R-134, will exist in the vapor phase
at standard temperature and pressure (STP). However HFE-7000 will not boil at atmospheric
pressure until 96.5°F. The pressure-enthalpy diagram is shown in Figure 2 with lines of
constant temperature, specific volume, and entropy. The positive slope of the saturated vapor
curve can be clearly seen, placing the fluid definitively in the “dry” category – as the fluid
proceeds through the expander and experiences a sudden reduction in pressure, the fluid
should not condense.

Figure 2. P-H diagram for HFE-7000.

The state analysis used to calculate the states of the ORC was created in a spreadsheet
and is semi-automated. Certain inputs are defined and the spreadsheet calculates all the
states by looking up relevant values in HFE-7000 property tables. With the automated
spreadsheet theoretical goals can be quickly achieved by adjusting inputs. For example, to
10

achieve a higher thermal efficiency, it is easy to see precisely what input temperature would
be required, or what expander efficiency, or what combination of these and other factors.
In the model, the following inputs are required to fully define the system:
 State 1 pressure
 Thermal sink temperature
 Pump head
 Pump efficiency
 Energy input (defined as thermal reservoir temperature in, temperature out, and flow
rate)
 Expander efficiency
In order to define these inputs and to improve the accuracy of the model, the
following assumptions were made (some of these values will be validated or corrected with
the system experiment):
 The pressure at State 1 (p1) is ambient – 101kPa.
 The thermal sink temperature is ambient (25°C), and the reservoir is large enough to
assume it is isothermal throughout the experiment.
 The flow rate of the working fluid is 0.38 L/s (6 GPM) – pump head and efficiency
are obtained from the performance curve in Figure 20.
 Thermal reservoir temperature is 100°C, with a drop in temperature of 14°C at 1 kg/s
flow rate – this provides approximately 60kW of power.
 Expander efficiency is 50% - this is based on isentropic efficiencies reported for the
Copeland ZR108KC-TFD compressor that range from 42.7% to 72.7% [see
appendix].
 There is a 5% pressure drop through the boiler and the condenser.
 There is a 6°C difference in the heat exchangers between the working fluid and the
reservoir fluid.
 Work done by the pump and work done through the expander are both isentropic.
For each state, only two properties are required. The properties used, and the
methods used to obtain them, are described below.
State 1 – Condenser to Pump – Pressure is assumed to be atmospheric, and
temperature is the same as the assumed condenser temperature plus an offset.
(1)
11

(2)

State 2 – Pump to Boiler – Since it is isentropic work, the entropy at state 2 is the
same as at state 1. Pressure is found by adding head obtained from the pump curve (Figure
20) to pressure in State 1.
(3)

(4)

State 3 – Boiler to Expander – Enthalpy (H) is calculated from the heat addition and
fluid flow rate. Pressure is assumed to drop 5% from P2.
(5)

(6)

State 4 – Expander to Condenser – Expansion is treated as isentropic, so entropy at


State 4 is the same as that at State 3. Pressure is assumed to drop to 5% greater than the
pressure after the condenser (P1).
(7)

(8)

Efficiency of the expander and pump have not yet entered into the calculations, as
they do not have a direct bearing on the working fluid. Rather, the overall system efficiency
depends on the efficiency of both components:

, where (9)

, and (10)

. (11)

Where ηe and ηp represent the efficiency of the expander and pump, respectively.
State calculations using the above techniques, assumptions, and component characteristics,
together with data provided by 3M can be found inTable 1.
12

Table 1. Theoretical State Calculations


State 1 - After Heat Rejection
p1 103 kPa TH2O 298 K
T1 304 K mdotH2O 1.00 kg/s
h1 67.735 kJ/kg
s1 52 J/molK
State 2 - After Pump
Head 300 kPa
Efficiency 0.2
p2 403 kPa
h2 67.83 kJ/kg
T2 304 K
State 3 - After Heat Addition
p3 382.85 kPa H2OT1 374 K
h3 269.7 kJ/kg mdotH20 1 kg/s
T3 368 K T2 360 K
s3 170.24 J/molK Qin 58.52 kW
State 4 - After Turbine
p4 108.42 kPa
Efficiency 0.5
h4 250.0 kJ/kg
T4 342 K

Analysis
Power 2.863 kW
Efficiency 4.67%
mdot 0.29 kg/s

Values in the table that are bold and highlighted yellow represent inputs; all other
cells represent calculated values. The model predicts that the system will operate at a
thermal efficiency of nearly 5% while producing 2.9kW of electric power.
This analysis is intended to lend confidence to the system testing as well as to provide
theoretical goals for each state. Knowing the ideal fluid states at various points in the system
should take the guesswork out of the testing. These points can be validated using
thermocouples and pressure gauges mounted on the system. This data will then be
incorporated into the model to increase its fidelity. The pressure and enthalpy of theoretical
system states overlain on the saturation curve of HFE-7000 can be seen in Figure 3.
13

Figure 3. P-H diagram of theoretical ORC states.


14

CHAPTER 3

SYSTEM DEVELOPMENT

The vast majority of work that was done in this project went into modifying and
troubleshooting the existing Rankine Cycle system to bring it to an operable condition.
Damage to the system and components that were not appropriate to the fluid under
investigation had to be overcome in order to perform an actual test. The system at the outset
of the project can be seen in Figure 4.

Figure 4. ORC system at outset of project.

Components that made up the system are as follows:


 3 Flatplate heat exchangers
 Condenser – Flatplate model C35
 Boiler – Flatplate model CH50
15

 Pre-heater – Flatplate model FP5X12-100


 Centrifugal pump – Magnatex MPT201
 Air motor functioning as expander – N/A
 Electrical generator – N/A
 Oil separator – N/A
 Oiler – N/A
 Reservoir tank – One gallon pressure tank
 Various valves, pressure gauges, and thermocouple wells
 2 charge ports

HEAT SOURCE
Finding an appropriate heat source to provide energy to the ORC was critical to the
project. Any adequate heat sources must be near 100°C, measurable, and safe. In addition, it
should be controllable and large enough to power the system for several minutes. Several
options were investigated: steam, engine waste heat, and water.

Steam
Steam was an obvious candidate initially as it could be easily regulated, it was readily
available in the building heating system, and it supplied the temperatures required by the
ORC. However, it became apparent that the amount of steam required to run the system at
the desired power level was more than could be spared. Steam also presented a potential
safety hazard. For these reasons steam was abandoned as a potential energy source.

Internal Combustion Engine


The majority of the experiments performed on the ORC utilized the waste heat from
an internal combustion engine as an energy source. The primary benefit of this heat source
was that unlike the steam, it offered total control over when heat could be provided.
Drawbacks to the automobile engine heat source included excessive noise and exhaust
fumes, but the primary limitation was the lack of control over the quantity and quality of heat
that the engine was capable of producing.
Initially the waste heat was harvested from the exhaust by diffusing it through an air-
to-liquid heat exchanger (in this case, an automobile radiator). However, this method did not
16

provide enough energy to run the system at the desired levels. Ultimately, the heating fluid
line was linked directly with the cooling system of the engine with an in-line pump. In this
way, energy was removed from the engine block as the cylinders were firing and pumped
directly into the heat exchanger to heat and boil the working fluid.
This setup was successful in that it managed to raise the heating fluid to above 180°F,
though not any higher. However, there were several critical shortcomings that made the heat
source unacceptable. First, the flow rate was unknown – because of the thermostat in the
engine, the coolant pump would turn on and off depending on the amount of energy removed
by the ORC. Although the in-line pump was continually circulating, the flow rate would
vary depending on the state of the engine’s pump. Also, the engine would frequently
overheat, causing the pressure relief valve to open on the radiator, adding further variation to
the already erratic flow rate. Second, temperature was unknown – the thermometer on the
engine would fluctuate, requiring temperature measurement at the inlet of the boiler heat
exchanger, which is an inaccurate one. Lastly, energy input was unknown – the engine was
manually throttled up and down to provide more or less energy to the ORC, but there was no
way to quantify the energy.
Because the use of the engine as an energy source made the accurate gathering of data
impossible, it was abandoned in favor of a more controlled solution.

Water Heater Source


In the end a water heater was chosen as an energy source for the ORC. The water
heater offers the ability to easily control temperature, monitor the amount of energy entering
the system, and maintain a consistent flow rate. A water heater had been previously
considered but it was thought that safety features would prevent it from achieving the
temperatures that the system required. In fact, the temperature is only limited by the boiling
point of the fluid, which is water in this case. The water heater selected has a 150 L (40
gallon) tank and two electric heating elements that each draw approximately 3600 Watts.
The primary shortcoming of the water heater is its limited energy capacity – if the system
produces 2 kW at 5% thermal efficiency, it would require 40kW of power input from the
water heater. Rejecting energy at this rate, even with the heating elements powered, the 150
liters of water would lose about 3°C/minute, which would reduce the water to a temperature
17

below what is required to boil the HFE-7000 at the operating pressure within 6 minutes.
Further, because of inefficiencies in the heat exchanger, the water must be several degrees
warmer than the working fluid. Without multiple tanks, experiments on the ORC can only
last 2 or 3 minutes before the heat source is exhausted.
Energy is transferred from the water tank to the heat exchanger using a high-
temperature circulator pump. After a 1/12 HP pump failed to achieve adequate flow, a 1/8
HP pump was installed. The larger pump is capable of generating a flow rate of
approximately 1kg/s in the ORC system. All tubing between the tank and heat exchanger is
insulated. The water tank with pump and insulated tubing is shown in Figure 5.

Figure 5. Heat source installed.


18

HEAT REJECTION SINK


The theoretical analysis performed above shows that the ORC achieves an overall
thermal efficiency of less than 5%, which means that 95% of the energy put into the system
(minus minimal losses) needs to be removed at the condenser. The simplest approach to
remove this amount of energy is with a large thermal sink. In this case, a plastic tub was
filled with 340 L (90 gallons) of water, and a ¼ horsepower sump pump was utilized to
circulate the water through the condenser heat exchanger (see Figure 6). This approach
works well at first because the temperature of the water is low enough to condense the HFE-
7000. However, after the rejected heat warms the thermal sink, the working fluid is
condensed less efficiently. In the previous example of a 2kW system with 5% efficiency, the
tub warms at a rate of approximately 1.6°C/minute. In order to maximize system efficiency,
the sink must be maintained at a low temperature. This can be accomplished with ice, with
frequent replacement of the water, or by leaving long intervals between tests for the bath to
cool to the ambient temperature.

Figure 6. Heat sink installed.


19

SYSTEM COMPONENTS
The components that came with the original system were for the most part well-sized
for use with HFE-7000. Specifically the pump and the air motor were capable of delivering
the amount of power desired from the modified system. The heat exchangers were
sufficiently large to transfer enough energy to and from the working fluid to operate the
system at the desired power levels. The original system also included a large oil separator
immediately downstream of, and directly above, the expander. The majority of the time
spent on the project left the system as-is, simply connecting heating- and cooling-fluid
circuits, charging it with HFE-7000, and measuring power output. However, due to several
factors, no output was ever achieved. First, the lack of an adequate heat source made
delivery of sufficient energy to the system impossible. Second, the expander had a
substantial leak that was either caused or exacerbated by exposure to HFE-7000. Lastly, the
oil expander acted as a large thermal sink, causing the fluid to condense and fall back into the
expander, inhibiting its function. The last two factors necessitated a replacement or repair of
the expander and a reconfiguration of the system.

Heat Exchangers
The heat exchangers are comprised of an evaporator rated to 220 kW (50 tons) at 7.6
L/s (120 gpm) and a condenser rated to 154 kW (35 tons) at 6.6 L/s (105 gpm). The actual
flow rate through the boiler and condenser is approximately 1 L/s, yielding a capacity of
37kW and 29kW, respectively. The heat exchangers are well-sized for this system which
absorbs 30kW and rejects approximately 29kW. All heat exchanger types are flat plate type
which require less volume for equivalent capacities.

Pump
Based on its pump performance curve (Figure 23, p. 56), the Magnatex MPT201 is
able to provide the appropriate amount of head at the desired flow rates. Further, the pump
housing is magnetically coupled, which requires no seals around the shaft and reduces the
possibility of leaks.
20

Expander
A variety of expander types are available for this particular application, most notably
the multi-vane expander, screw expander, scroll expander, and micro-turbine. Each has
advantages and drawbacks that must be considered before any one is selected.

MULTI-VANE EXPANDER
The multi-vane expander (MVE) is a positive displacement expansion device
invented by Charles C. Barnes in 1874, originally intended as a compressor. It utilizes a
rotor turning eccentrically inside a cavity. Vanes on the rotor maintain contact with the
cavity walls (see Figure 7). Advantages to the MVE include flat operating efficiency curves
over a wide range of conditions, low speeds (approximately 3000rpm) that can match
generator speeds without a gearbox, the ability to operate in the presence of liquids and wet
vapors, minimal maintenance with little lubrication requirement, and proven operation with
organic working fluids (O’Callaghan et al., 1985).

Figure 7. Multi-vane expander.

The authors performed an extensive analysis of two different varieties of MVE


(circular and non-circular), taking into account the major modes of losses such as breathing,
internal leakage, and friction. The non-circular MVE was a converted commercially
available refrigeration compressor. The circular MVE was a prototype designed by the
authors in which the stator cylinder comprised two circular arcs, one acting as a sealing arc.
21

The theoretical simulation the authors performed was compared to experimental data
gathered using a low-grade energy ORC with R-113 as a working fluid. The fluid was mixed
with 5% oil by weight to keep the expander lubricated during operation.
A number of factors were modeled and measured, including flow rate, rotational
speed, cell pressure, torque, mechanical efficiency, power output, and isentropic efficiency.
The theoretical model often aligned with experimental results with errors of less than 10%.
Testing was limited by power available at the boiler – 85kW at approximately 125°C.
While mechanical efficiency was measured at over 80%, the isentropic efficiency of either
type of MVE never exceeded 60%. This value corresponded to the peak power point and
occurred at a rotational speed of 3200 RPM.
Work has continued over the years with the investigation and implementation of
MVEs as prime movers in low-temperature ORCs due to their wide availability, reliable
performance over a range of conditions, and compatibility with a variety of fluid types. A
multi-vane expander would be a reasonable choice for use in the present research.

SCREW EXPANDER
The helical screw expander is likewise a positive displacement device. It is
comprised of two meshing, helical, counter-rotating screws (see Figure 8). As the fluid
expands and moves along the screws, it drives (rotates) them apart to increase volume. This
work is geared to a single shaft which can be used to power a generator.
Since the 1970s two-phase screw expanders have been investigated and utilized for
the recovery of power from liquid or low-dryness geothermal brines by expanding steam or
even pressurized hot water. Smith et al. (1999) worked to use the two-phase expansion
process with organic working fluids for which exit volume flow rates and expansion ratios
are significantly reduces for similar heat utilization at certain temperatures.
The authors achieved an expander efficiency of greater than 75% in an organic
Rankine cycle using R-113 as a working fluid. Fundamental improvements to the expander
design were required to achieve this high efficiency.
Like all other positive displacement devices, the seal is critical to prevent internal
leakage. In order to prevent direct contact but also achieve a seal between the lobes of each
rotor in a screw expander, two methods of lubrication are device types have been developed:
22

Figure 8. Screw expansion device (oil injected). Source: Smith, I.


Stosic, N., & Kovacevic, A. (1999). Power recovery from low cost
two-phase expanders. Centre for Positive Displacement
Compressor Technology, City University, London.

oil injected and oil free. The oil injected machine, which relies on oil carried by the
compressed gas in order to lubricate the rotor motion, seals the gaps and reduce the
temperature rise during compression without requiring internal seals. The oil injected type of
machine is simple in mechanical design, cheap to manufacture, highly efficient, and widely
used as a compressor. The oil free machine separates any oil from the working fluid by
preventing contact between the rotors with lubricated meshing timing gears outside the
working chamber. Internal seals are required against the bearings and chamber walls. These
additional parts and requirements cause the oil free machine to be considerably more
expensive than its oil injected counterpart. Smith et al. found that the amount of oil required
for the oil injected machine could not be carried by the fluid vapor, but that the standard oil
free machine was cost prohibitive and so were forced to develop an alternative.
Using a new lobe profile that nearly eliminated any sliding motion between rotors in
favor of rolling motion, and reduced contact stress, significantly reduced the need for oil to
the point where a small amount of injected oil could be used. Further, this new design did
not require a timing gear as a typical oil free design would. These improvements resulted in
23

a relatively inexpensive expander device that was compatible with organic working fluids
that can carry only a small amount of oil, yet highly efficient.
Screw expanders depend on precise numerically-controlled machining to achieve a
leak-resistance fit, especially with oil free types. Because of the tight sealing requirements,
they tend to work better with wet fluids. Because the authors worked with a two-phase
system, their results that show high efficiency will not likely transfer to the dry system under
investigation in this paper. A dry fluid system would still require seals that greatly increase
the cost of the machine.

SCROLL EXPANDER
The scroll expander is the last type of positive displacement device that was
considered in this study. The device is comprised of a stator and a rotor scroll (resembling
spirals). The rotor orbits the stator eccentrically without rotating, trapping pockets of fluid
against the stator. As the rotator processes against the stator, this volume of trapped fluid
expands (see Figure 9). In reality, the fluid itself is under pressure, so its expansion drives
the rotation of the rotor, creating useful work. The scrolls and shaft are designed to minimize
the gap between rotor and stator as this is a source of leakage.

Figure 9. Scroll expander operation.

The scroll expander is an inexpensive device due to its widespread use and low-
complexity. Scroll compressors, which operate on the same principle but in reverse, are
commonly used in both automotive and building air conditioning systems.
Several recent studies have been performed characterizing myriad aspects of the
scroll expander.
24

Lemort et al. (2009) present the results of an experimental study carried out on a
prototype of an open-drive oil-free scroll expander integrated into an ORC working with
refrigerant HCFC-123. The authors identify eight parameters of a scroll expander for semi-
empirical modeling. Then, relevant parameters are such as the mass flow rate, the delivered
shaft power and the discharge temperature are determined. Also, secondary variables such as
the supply heating-up, the exhaust cooling-down, the ambient losses, the internal leakage and
the mechanical losses were obtained. The maximum deviation between the predictions by the
model and the measurements was 2% for the mass flow rate, 5% for the shaft power and 3 K
for the discharge temperature. The validated model of the expander was used to quantify the
various losses and to introduce possible design alterations to achieve better performances. It
was concluded that internal leakages and, to a lesser extent, the supply pressure drop and the
mechanical losses are the main losses affecting the performance of the expander.
Lemort and Quoilin (2009) have recently evaluated the efficiency of scroll expanders in
organic and inorganic Rankine cycles with promising results. Between two studies, over 40
steady state performance points were gathered for R123 and water evaluating the effects of
boiler temperature, mass flow rate, condenser temperature, and expander rotational speed on
shaft power, the expander isentropic effectiveness, and the cycle efficiency. Expander
efficiencies ranging from 42% to 68% were observed yielding system efficiency up to 7.4%.
Loss analysis of scroll expanders using R134a refrigerant was investigated by
Shigemi, Toshio, and Akira (2009). A method was proposed to estimate losses using P-V
diagram. The authors suggest that leakage in suction process was large, so tip seals of scroll
clearance are very important to obtain high efficiency. A new suction port system was
designed which decreased the leak by about 10%, and in turn increased expander efficiency
by about 10%. Other relevant results were also obtained by the authors, such as the effect on
efficiency of pressure ratio, mass flow rate difference between measured and theoretical
values versus the increase in the rotational speed and the decrease in the pressure ratio of
expander, and maximum shaft power and expander efficiency. An expander efficiency of
74% was obtained at the pressure ratio of 3.7 and rotational speed of 50 Hz.
Mathias, Johnston, Cao, Priedeman, and Christensen (2009) tested application of
gerotor and scroll expanders to determine their applicability in producing power from low-
grade energy using R123. Their experimental study showed that both types of expanders
25

were good candidates to be used in an ORC. The experiment produced 2.07 kW and 2.96
kW, and had isentropic efficiencies of 0.85 and 0.83, respectively. The authors also
presented results incorporating some strategies to improve cycle efficiency. One such change
pertains to matching the flow rate of the working fluid with the inlet pocket volume and
rotational speed of the expander. The volumetric expansion ratio of the expander was also
matched to the specific volume ratio of the working fluid across the expander. The network
was calculated after subtracting the power required by the pump and the condenser fan –
which gave 6271 W of net power at an overall energy efficiency of 7.7%.
Scroll devices are readily available and inexpensive, partly due to their widespread
use as air conditioning compressors. Scrolls customized for use as expanders are also
commercially available. A 3kW oil-free scroll expander from Air Squared costs $3,300 and
can achieve an isentropic efficiency of over 80% (http://www.airsquared.com).

TURBO EXPANDER
The micro turbine, or turbo-expander, is the only non-positive displacement device
considered. It operates in a similar manner to the expansion portion of a gas turbine engine
(combustion and compression removed). A high pressure gas is directed past turbine blades
causing them to rotate as the gas expands. This rotational work can be harvested for
electrical power. The rotor can be an axial or centrifugal type, with the centrifugal being
more common because of its frequent use in automotive turbochargers. The turbo-expander
is standard for large-scale Rankine cycles due to its high efficiency. In small-scale
applications, drawbacks to the turbo-expander include high cost, high shaft-speed that must
be geared down to work with an electric generator, and a low tolerance for varying gas inlet
pressures. One such experiment in which a turbo expander was used in a low-power
application was performed by Teagan and Clay (1973). Their research investigates the
performance of a 3 kW expander and generator set driven by an organic Rankine cycle. The
turbine speed of 70,200 rpm was geared down to 3,600 RPM. This gearing was done
internally to the sealed system so that a low speed shaft seal with proven performance could
be used to transmit the turbine output to the generator. Fluorinol-85 was used as the working
fluid. The conditions at the inlet to the turbine were 530F and 480 psi. At the exit of the
turbine conditions were 202F and 29 psi.
26

The operating pressure and temperature of the system investigated above are both
higher than can be achieved in this study. No documented uses of the turbo expander at
pressures and power levels more appropriate for this study have been found. Because of this
poor heritage, and because of its high cost, the turbine expander is a decidedly poor choice
for the present research.

EXPANDER SELECTION
After careful consideration of available expansion devices, a scroll-type device was
chosen as a replacement for the air motor. The scroll device has several advantages: few
moving parts, wide availability, relatively low cost, and proven performance with several
common working fluids. Significantly, extensive research has been performed with scroll
devices in similar systems. These studies form a strong basis for continued work with the
scroll expansion device, lending confidence to the choice and helping to avoid pitfalls as the
device is optimized for this particular system. A qualitative comparison of the different
expander types can be seen in Table 2. The quantity, applicability, and recency of research
focusing on each expander type, as well as any commercial use, define the “heritage” factor.

Table 2. Expander Comparison Summary

Type  Performance Cost  Heritage 


Rotary Vane  +  $  + 
Screw  ‐  $$  + 
Scroll  ++  $  ++ 
Turbine  ++  $$$  ‐‐ 

The scroll device chosen for this project is actually a compressor designed for air
conditioning systems – model ZR108KC-TFD made by Copeland (see Figure 10)
[specifications in appendix]. Conversion of the Copeland compressor, though not ideal
because it is not optimized for this use, has been successfully accomplished before (Ingley,
Reed, & Goswami, 2004; Kim et al., 2007). As a compressor, this device is rated at 9
horsepower, or 6.7 kW, which is more than adequate for the amount of power expected from
the ORC. It was important to allow flexibility in the system in case more or less power was
27

Figure 10. Expander schematic.

desired; the expander selection was intentionally oversized with the knowledge that lower
power operation would cause a reduction in efficiency.

COMPRESSOR CONVERSION
Since the expander is actually a compressor operating in reverse, substantial
modifications were required in order to ensure both its operation as an expansion device and
its integration in the ORC system. Before any modification began, the device was tested
28

under normal operation (with 3-phase power supplied) to verify its functionality. Then
compressed air was injected into the outlet (which became the inlet of the expander) to test
its operation in reverse. It was quickly verified that device check valve was preventing
reverse flow, and the case had to be opened (see Figure 11). The scroll device is shown in
Figure 12.

Figure 11. Expander with top removed and check valve assembly.

Figure 12. Scroll device sides A and B.


29

After removal of the check valve, an injection of compressed air directly into the
scroll device verified its operation in reverse. However another function of the assembly was
to route fluid from the scroll outlet to the case outlet. The removal of the check valve
allowed gas to move freely from the inlet to the exit of the case, completely bypassing the
scroll device.
In the compressor, fluid enters at the inlet then is free to diffuse throughout the case
before it is compressed through the scroll and routed through the check valve and toward the
outlet at the top of the case (see Figure 13). It would have been possible to modify the check
valve assembly to allow reverse flow and still maintain proper fluid routing, but three factors
made this an undesirable solution: head loss through the modified valve would decrease
system efficiency, the tolerances that were required to achieve a seal when the cap was
reattached to the case were extremely tight, and using the port lower on the case made system
integration difficult.

Figure 13. Compressor flow diagram.


30

Instead of keeping the existing fluid routing, it was decided to route the incoming
fluid directly through the top of the cap into the scroll, and then allow the fluid to diffuse
through the case after expansion and exit through the original exit port (see Figure 14).

Figure 14. Expander flow diagram.

The modification to the fluid routing through the expander required widening the inlet
of the scroll device (see Figure 15) in order to accept the 1.5” pipe diameter used elsewhere
in the system. The inlet was widened further to accommodate o-rings used for a seal.
The previous inlet port was capped since it was no longer needed. The previous
outlet port was widened to accommodate 1.5” tubing. The new inlet and exit tubing was
welded to the cap. A flange was also welded around both the cap and the case in order to
facilitate removal of the cap during troubleshooting. For the same reason, unions were
installed on either side of the expander to allow easy removal (see Figure 16 for cap
modifications).
31

Figure 15. Scroll device installed with inlet machined.

Figure 16. Cap before and after modifications.

PLUMBING CONFIGURATION
The removal of the oil expander opened the window for additional modifications to
the system plumbing. First, the expander was moved to the highest point possible in the
system in order to eliminate the risk of liquid HFE-7000 falling back into the expander as it
did in the original configuration. Figure 17 shows the expander installed in the system.
32

Figure 17. Expander installed in ORC system.

Second, extraneous piping was removed in order to minimize parasitic thermal losses.
This involved lowering the condenser which required removal of one charge port, and
removing stress-relief tubing. The stress relief tubing is a potentially beneficial feature for a
system undergoing deep thermal cycles, however it was decided that the low operating
temperature of the system made stress-relief unnecessary. Lastly, the line between the pump
and the boiler was straightened (see Figure 18), which reduced head loss through this high-
pressure line and also allowed a charge/drain port to be installed at the bottom of the system.
Positioning the drain port at the lowest point is when HFE-7000 is used because it is a liquid
at room temperature and cannot be removed like traditional working fluids.
33

Pre heater

From pump

Figure 18. Pump to pre-heater link before and after modification.

ELECTRICAL LOAD
The air conditioning compressor that was converted to an expander uses a 3-phase
induction motor to power the scroll. This complicated the conversion of the compressor
because induction motors utilize current in the stator to induce a magnetic field in the rotor
which is used to generate torque. When an induction motor is used as a generator, there is no
current provided from an external source, and so no magnetic field can be induced in the
rotor. However, if capacitors are wired between stator leads and a magnetic field can be
induced somehow, even for a short period, the generator will become self-sustaining. There
are two ways to apply a current to the stator – an external power source, such as a car battery,
can be used to “jump start” the coils, or the tiny amount of residual magnetism that almost
always remains in the copper windings can be exploited to induce small currents which in
turn induce more magnetism until the coils build up enough charge to operate at maximum
voltage. When using the latter method, which was chosen for this project, it is critical that no
load be applied while charge is building, since virtually any load would be too large for these
tiny seed currents to be produced.
Motor run capacitors of 35 microfarad rated at 370VAC were selected (see Figure
19). These are more capacitive than is strictly required, but it is safer to err on the high side
as larger capacitors will not inhibit the charging in any way, but undersized capacitors could
short.
34

Figure 19. Run capacitor wiring.

Various loads were considered that are capable of dissipating multiple kilowatts of
power. Ideally, the generator would be wired to the hot water heater to dissipate any power
produced directly back into the heat source. This was tried without success – unbalanced
loads disable 3-phase generators and only 2 heating elements were available in the water
heater. If a 3rd element were available, a resistive load like this would have ample capacity to
dissipate whatever power the generator produces. Also, resistive loads have no reactive
element, so all the power measured is real power with no imaginary component that must be
factored into calculations later. Ultimately a 3-phase motor was selected as the load;
although the reactive element of an electric motor requires some calculation to find the total
effective power, a motor offers the benefit of guaranteed functionality with a 3-phase source.
Switches were also wired along 2 of the lines to allow manual application of the load once
sufficient charge is achieve in the induction generator. No third switch is required because a
35

single line does not complete any circuit through the motor and thus cannot draw any power
from the generator. The electric connection between the generator and the load is shown in
Figures 20 and 21.

Figure 20. Electrical setup diagram.

Figure 21. 3-phase load with clamp-on


ammeter and switches.
36

MEASUREMENT
The ORC is equipped with thermocouple wells at critical locations throughout the
system, allowing measurement of the working fluid temperature. The system also has
pressure gauges mounted in critical locations, but during system charging it was discovered
that the offset and/or accuracy of these gauges are poor enough to add considerable
uncertainty to system measurements. Though correction factors were added during this
impromptu calibration, data from the pressure gauges are suspect and treated as primarily
qualitative.
Measurement of the heating and cooling fluids is critical for calculating the total
energy added to and removed from the system. Measuring temperatures at the inlet and exit
of the heat exchangers was done by using a thermocouple and a heat gun to measure the
temperature of the fittings on the heat exchanger. This method proved to be inaccurate and
was abandoned in favor of a time-averaged bulk reservoir temperature measurement. During
system testing, temperatures of the thermal reservoir and sink were recorded at fixed time
intervals. Data were also taken in a similar way without the system running in order to
account for error caused by thermal losses from the reservoirs. It turned out that both the
reservoir and sink lost a negligible amount of energy over the time frame of the experiment –
due to effective insulation and relatively large thermal mass.
Measurement of the electrical output is done using a clamp-on ammeter around one
of the three wires powering the 3-phase motor and a voltmeter probing across one of the
capacitors.

WORKING FLUID
Many factors must be considered in the selection of the working fluid. Besides
thermodynamic performance, these also include environmental impact, safety, cost, and
availability.

PERFORMANCE
The performance of refrigerants in Organic Rankine cycles has been exhaustively
researched. One common characteristic considered in fluid selection is the slope of the
37

saturation vapor curve. Because of the nature of the low-quality heat source, the superheated
approach employed by a traditional Rankine cycle cannot be utilized; the fluid will be close
to its saturated vapor curve as it enters the expander. This means that wet fluids will
condense as they lose enthalpy through the expander, an undesirable characteristic.
Conversely, dry fluids become superheated as pressure drops through the expander. This
superheated gas can be used as a regenerator to pre-heat condensed fluid. Hung et al.
recently found that a wet working fluid with a steep curve outperformed dry fluids (Hung et
al., 1997), despite significant work that has been done to investigate dry fluids (Hung, 2001;
Hung et al., 1997; Mago et al., 2008). However no regenerator was implemented in the work
done by Hung. Isentropic fluids, those with curves that are nearly vertical, seem to perform
best in a non-regenerator system (Quoilin, 2007).
Another factor that affects performance of a fluid is its heat of vaporization, or latent
heat. Denser fluids with greater latent heat can move more energy through the system with
less mass flow. This allows a reduction in the size of all components and an increase in
thermal efficiency due to lower losses through large heat exchangers and a larger pump.
Lastly, boiling point of the fluid is a critical consideration since the intended use is
low-grade heat recovery. The fluid must be able to undergo a phase change at a reasonable
pressure (approximately 6 atmospheres) at the intended operating temperatures (below
100°C ). This particular criterion narrowed down the available refrigerants significantly.

ENVIRONMENTAL IMPACT
Environmental impact of a working fluid is primarily characterized by three
parameters: ozone depletion potential (ODP), global warming potential (GWP), and
atmospheric lifetime (ALT). ODP is a measure of the relative capacity for ozone depletion
compared to trichlorofluoromethane (CFC-11), which has been assigned an ODP value of
1.0. Traditional refrigerants including CFCs and HCFCs have ODPs between .1 and 1.
Many modern refrigerants, such as hydrofluorocarbons (HFC) and hydrofluoroethers (HFEs)
have eliminated chlorine from their chemical structure due to the high reactivity of chlorine
with ozone, and so have ODPs of zero. GWP is a factor that uses a similar method to
estimates how much the chemical will contribute to global warming over a fixed time
period. In the case of GWP, the reference chemical which has been given a GWP value of
38

1.0 is carbon dioxide. The factor is based on the total amount of infrared absorption, the
particular wavelengths of absorption, and the atmospheric lifetime of the chemical. location
of radiative factor, and decay rate of a chemical (Environmental Protection Agency [EPA],
2010). The ALT of a greenhouse gas refers to the approximate amount of time it would take
for human-caused contribution to an atmospheric pollutant concentration to return to its
natural level as a result of either being converted to another chemical compound or being
taken out of the atmosphere via a sink. This time depends on the pollutant's sources and sinks
as well as its reactivity. The lifetime of a pollutant is often considered in conjunction with the
mixing of pollutants in the atmosphere; a long lifetime will allow the pollutant to mix
throughout the atmosphere. Average lifetimes can vary from about a week (sulfate aerosols)
to more than a century (chlorofluorocarbons, carbon dioxide) (EPA, 2010).
For nearly 100 years CFCs have been used as refrigerants because of their desirable
performance and safety characteristics. Due to their high ODP they are being temporarily
phased out by HCFCs which have a lower ODP. However, even HCFC contribute to ozone
depletion to some extend and are being phased out as well (http://www.arap.org). The
Montreal Protocol of 1987 eliminates CFC production by 2010 (except for essential uses)
and HCFC production by 2030 and 2040 in developed and developing countries,
respectively.
Alternatives to HCFCs include HFCs and HFEs. HFCs provide many of the required
performance characteristics of CFCs and HCFCs and have zero ODP. As a result, they are
frequently used as a drop-in replacement fluid. However HFCs have a significant GWP so
have been identified by the Kyoto Protocol as a greenhouse gas and thus targeted for
reduction (http://www.arap.org).
HFEs show promise as an ultimate solution both to GWP and ODP concerns. HFEs
have zero ODP as well as very low GWP and ALT compared to HFCs. These properties
make HFEs desirable for use as a working fluid in energy conversion systems, but very little
investigation has been previously performed into the use of HFE fluids in such an
application.
A comparison of the three factors discussed for several commonly used refrigerants,
and the fluid under investigation, is shown in Table 3.
39

Table 3. Environmental Factors for


Possible Working Fluids
   ALT ODP GWP
CFC‐11  45  1  3660 
HCFC‐22  12  0.034  1710 
HCFC‐123 1.3 0.012 53
HFC‐134a 14 0 1320
HFE‐7000 4.9 0 370

SAFETY
The American Society for Heating, Refrigeration and Air Conditioning Engineers
(ASHRAE) defines the level of toxicity and flammability of a refrigerant as one of 2 and 3
classes, respectively. Toxicity can either be class A if the refrigerant has not been identified
as toxic at concentrations less than or equal to 400 ppm by volume, or class B if it has been.
The flammability of a refrigerant can be class 1 if there is no flame propagation in air at 21°C
and 1 atm, class 2 if its lower flammability limit (LFL) is greater than 0.10 kg/m3 at 21°C
and 1 atm and its heat of combustion is less than 19,000 kJ/kg, or class 3 if it is highly
flammable as defined by LFL less than or equal to 0.10 kg/m3 at 21°C and 1 atm or its heat
of combustion is greater than or equal to 19,000 kJ/kg. These two classes are written
together as a two-character figure (for example: A2 or B3).
The ASHRAE safety figure does not factor into the performance of the ORC, but is
nonetheless important to consider in the selection of a working fluid as it affects operator and
facility safety.

FLUID SELECTION
Due to largely environmental considerations, this study narrowed the list of possible
refrigerant candidates to HFEs. Within that category, boiling point was a discriminating
requirement that further narrowed the field.
After consideration of all the relevant factors that have been identified, HFE-7000
was selected as the working fluid for this project. HFE-7000 is a hydrofluoroether
manufactured by 3M (under the trade name Novec 7000). It is a segregated HFE, meaning it
has no hydrogen atoms within the molecule bonded directly to fluorine, but instead has all
40

hydrogen-carbon bonds segregated from fluorine by ether oxygen, reducing its atmospheric
lifetime (Owens, 1998). It has a GWP of 370, zero ODP, and an ALT of 4 years. It is a safe
fluid with low toxicity and zero flammability (ASHRAE A1) (3M Electronics, 2005). From
a performance standpoint, it is a dry fluid of high-density, high latent heat, and a breakdown
temperature of 186°C, making it a potentially good replacement for current HCFC and HFC
fluids. Other benefits include a high heat of vaporization allowing system components with
lower pressure capability, and compatibility with a wide range of materials including brass,
copper, and a wide variety of plastics. Husband and Beyene (2008) have demonstrated the
feasibility of a low-grade heat recovery system that can produce electrical power using HFE-
7000 as the working fluid. A theoretical efficiency of 11% for a 10kW system was shown at
a pressure ratio of 6 and input temperature of 240°F.
41

CHAPTER 4

RESULTS AND DISCUSSION

EXPERIMENTAL DATA
The experiment performed with the ORC system successfully produced electric
power. The system experiment began with the water heater temperature at boiling and the
condenser tank at 80 deg F. Circulator pumps for both reservoirs were on to ensure the heat
exchangers were in equilibrium with the fluid and would not cause any transient effects. The
3-phase load was electrically detached from the generator using the two switches (see Figure
18). A timer was started when the compressor was turned on and valves were opened.
Almost immediately, the expander began spinning (can be felt and heard). Within seconds,
voltage across the capacitors built from less than 1V to nearly 300V. Once maximum
voltage was reached and the rotor was properly magnetized, the two switches were closed to
connect the load. Temperature data for the two water tanks and electrical data between the
generator and motor were taken at 30 second intervals.
The system produced 937 W electrical power using an input of 30 kW. This
corresponds to a gross thermal efficiency of 3.1%. Results are summarized in Table 4.

Table 4. Summary of Experimental Data


Qin Qout Wout
Volume 20700 in³ Current 2.6 A
40 89.6 gallons Voltage 208 V
151.2 338.7 liters Power 937 W
Mass 151200 338727 grams
ΔT ‐12.5 6.5 °F ηth 3.1%
Δt 180 180 s ηe 31%
ΔE 4390365 5114490 J
P 30444 27414 W
Δ 3030 W
42

The power measured at the output of 2.6A at 208V corresponds to the “real”
component of a single line. In order to calculate total power, “reactive” power must be
included as well. The theory behind the calculation is not within the scope of this paper, but
total power can be derived from real power by multiplying by the square root of 3 (Manhire,
2004).
It can be seen from the table that the amount of energy lost between the boiler and the
condenser (3030 W) is significantly higher than the amount of energy produced as electricity
(937 W). The discrepancy between these numbers primarily represents the efficiency of the
expander according to the following equation:
(12)

This equation assumes that other losses, such as thermal losses from the pipes and the
expander, are negligible; this is a safe assumption because the system is well insulated. From
the equation the efficiency of the expander is found to be 31%.

UPDATE TO THEORETICAL MODEL


The theoretical state model was updated with data gathered from the system
experiment. Data updated include condenser temperature, expander efficiency, and energy
addition to the boiler. Using energy addition to the boiler along with the enthalpy at states 2
and 3 allows the calculation of the mass flow rate within the system according to the
following equation:
(13)

The flow rate is calculated to be 0.15 kg/s. This can then be used, along with the
temperature at the pump (T1), to find the efficiency and head of the pump, which is also
updated in the model. The new state calculations can be found in Table 5. It should be noted
that a mass flow rate of 0.15 kg/s corresponds to a volumetric flow rate of slightly over 400
cubic feet per hour (CFH) at the inlet.
Also included in Table 5 is a calculation of the work done by the pump, and the
amount of electrical power (Win gross) required to provide this energy at the rated
43

Table 5. Experimental State Calculations


State 1 - After Heat Rejection
p1 103 kPa TH2O 302 K
T1 304 K mdotH2O 1.00 kg/s
h1 67.735 kJ/kg
s1 52 J/molK
State 2 - After Pump
Head 520 kPa Win(net) 0.02 kW
Efficiency 0.07 Win(gross) 0.33 kW
p2 623 kPa
h2 67.89 kJ/kg
T2 304 K
State 3 - After Heat Addition
p3 591.85 kPa H2OT1 372 K
h3 267.35 kJ/kg mdotH20 1 kg/s
T3 370 K T2 365 K
s3 166.14 J/molK Qin 30.43 kW
State 4 - After Turbine
p4 108.42 kPa
Efficiency 0.31
h4 242.4 kJ/kg
T4 334 K

Analysis
Power 1.180 kW
Efficiency 2.80%
mdot 0.15 kg/s

efficiency. The “efficiency” value near the bottom of the table is thermal efficiency, for easy
comparison to data reported in Table 4.
A chart of the states overlaid on a pressure-enthalpy (P-H) diagram of HFE-7000 can
be seen in Figure 22.

DATA COMPARISON
The data between experiment and theory agree fairly well, with the model predicting
higher power output and lower efficiency than was actually measured. In a well-insulated
system with well-characterized components, the model should accurately predict real-world
system performance. In this case, the power output matches within 20% and the efficiency
matches within 13%. This comparison serves to validate the state calculations and the fluid
property tables, but does recommend fine-tuning of the model for this particular system.
44

Figure 22. P-H diagram of actual ORC states.

More accurate system data would be required to determine precisely which


assumptions in the model lead to a discrepancy in the output characteristics. However, for a
system for which accurate component properties are known (such as head loss through heat
exchangers), future models can be optimized before any hardware is built.

DISCUSSION
A 3% thermal efficiency for this system is reasonable to expect; the Carnot cycle
efficiency at the experimental temperatures is only 18.8%. However, significant
improvements should be possible with further optimization of key components. A goal of
one-half the Carnot efficiency – 9.4% – is achievable with HFE-7000.
An increase in thermal efficiency can be achieved by addressing several significant
issues that arose after analyzing measured data from the system test and comparing the data
to the model. Those issues relate to the pump, the expander, the heat source, and the heat
sink.

Pump Performance
First, the pump was found to be oversized for the flow rate achieved (which is largely
dependent on the capacity of the heat source, to be discussed below). The pump efficiency
45

was found using the manufacturer performance curves (see Figure 23) to be 7% – an order of
magnitude lower than what should be expected for an application of this size. Even with a
greater flow rate, the pump can only achieve an efficiency of 20%.

Figure 23. Pump performance curve.

In the model, the low efficiency results in a work input of 330W (“Win(gross)” in
Table 5). This represents over a third of the power generated by the system and so reduces
the overall thermal efficiency by that amount. With a highly efficient pump – say, 90% – the
impact on thermal efficiency would be less than 1%.
In practice, the effect of the pump on system efficiency is even worse. The rated
power of the motor driving the pump is almost 1.44 kW – 0.5 kW higher than the power
output by the ORC. If the pump power were considered in the net thermal efficiency
calculation, the system would have negative efficiency. During operation, the pump feels
hot, implying additional friction or work that is not being applied toward pumping the fluid.
From both a theoretical and a practical standpoint, it is clear that the pump is not
appropriate for the system size and HFE-7000. It is oversized for the flow rate and head
required, and it is not designed for, and not necessarily compatible with, a fluid such as HFE-
46

7000 that is more dense and less viscous than water. For this application, a smaller
centrifugal pump or a positive displacement pump could achieve an efficiency greater than
75%.

Expander
The overall efficiency for the conversion of thermal to electric energy between the
boiler and condenser was found to be 31%. Some thermal losses must be expected through
the heat exchangers, the plumbing, and the case of the expander itself. However, the system
is well insulated, so it is unlikely that the conversion efficiency of the expander/generator is
higher than 35%.
Considering that the scroll device was able to achieve an efficiency of 72.7% [see
appendix] when operating as a compressor, a similar performance should be possible when
operating in reverse, so long as similar conditions are applied. Further, a compressor
converted for use as an expander has been demonstrated to operate at nearly 75% efficiency
(Kim et al., 2007). Two possible culprits that could account for the poor performance of the
scroll are pressure and flow rate.
An improvement to the performance of the scroll could be achieved with an increase
in pressure. However, since the pressure of the HFE-7000 is already its vapor pressure at
State 3, the pressure cannot be increased given the limitations of the heat source. That is, the
pressure of the fluid cannot rise unless its temperature rises as well, but the temperature is
already near the boiling point of water, the maximum available temperature with a hot water
heating system.
The volumetric flow rate of fluid through the expander is the other factor that strongly
influences the efficiency of the scroll. The flow rate of HFE-7000 was found to be 0.15 kg/s,
which corresponds to a volumetric rate of 400 CFH at the inlet to the scroll, and 2580 CFH at
its exit. Per the datasheet for the expander, rated displacement is 1060 CFH. This rated
value probably corresponds to the inlet when the scroll acts as a compressor, or the outlet for
the expander configuration. So rather than underpowering the scroll as it was originally
thought would happen, it is likely that the fluid is either overexpanding, or much of the
expansion is taking place in the expander case and the subsequent plumbing, rather than in
the scroll itself. Either of these scenarios would adversely affect the performance of the
47

scroll device because of leakage and friction. Further, the electric motor used to drive the
scroll when it acts as a compressor is optimized to operate most efficiently at a certain speed
(in this case, 3500RPM). Likewise, when acting as a generator, it operates most efficiently at
the same speed. The further the rotor speed varies from 3500RPM, either faster or slower,
the less efficiently the generator will operate. Both the scroll device losses and sub-optimal
generator performance likely account for the poor efficiency achieved during operation.

Heat Source and Sink


It became clear during testing that the thermal reservoir was not massive enough even
to adequately power the existing system. The 30kW of power that the system consumed
quickly decreased the temperature of the 150 L water heater (a 13°F drop in 3 minutes) even
with the heating elements powered. Likewise, the heat sink reservoir heated relatively
quickly. As the temperature of both source and sink change, the performance of the system
changes.
From Figure 22, it can be seen that the HFE-7000 is sub-cooled at State 1. This sub-
cooling decreases efficiency because it removes additional energy from the fluid that must be
added again at the boiler. An increase in condenser temperature so that the fluid at State 1 is
closer to the saturated liquid curve (37°C at p1) could increase overall system efficiency
appreciably.
Also from Figure 22 it is seen that the fluid is superheated at State 4. This excess
energy is currently wasted by being rejected to the heat sink. It could be better utilized in a
regenerator to contribute to the heating of the fluid between States 2 and 3.

Overall Efficiency Gains


If all of these improvements were to be made to the system: expander optimization,
pump optimization, and condenser temperature, system efficiency could reach nearly 10%
(see Table 6). The increase in efficiency over experimental results of approximately 7% are
made up primarily by expander optimization (5.5%), then by pump optimization (1%), and
lastly by condenser temperature optimization (.5%).
48

Table 6. Optimized State Calculations


State 1 - After Heat Rejection
p1 103 kPa TH2O 310 K
T1 312 K mdotH2O 1.00 kg/s
h1 76.985 kJ/kg
s1 58.26 J/molK
State 2 - After Pump
Head 520 kPa Win(net) 0.02 kW
Efficiency 0.75 Win(gross) 0.03 kW
p2 623 kPa
h2 77.13 kJ/kg
T2 312 K
State 3 - After Heat Addition
p3 591.85 kPa H2OT1 372 K
h3 267.35 kJ/kg mdotH20 1 kg/s
T3 370 K T2 365 K
s3 166.14 J/molK Qin 30.43 kW
State 4 - After Turbine
p4 108.42 kPa
Efficiency 0.75
h4 242.4 kJ/kg
T4 334 K

Analysis
Power 2.994 kW
Efficiency 9.74%
mdot 0.16 kg/s

HFE-7000
The use of HFE-7000 as a working fluid in a Rankine Cycle is clearly demonstrated
by this project. The fluid functions as other common refrigerants would with the added
benefits of being benign to human health and the environment. A comparison between HFE-
7000 and two other common refrigerants – R-22 and R-134 can be seen in Table 7. These
values are chosen to approximate States 3 and 4 of the system experiment, which is across
the expander, and to help understand the poor performance of the expander.
The ratios of pressure across the states are similar between all fluids, as are the
differences in enthalpy (not shown). One noteworthy difference is the relative density of
HFE-7000 – over 2 times higher than the other fluids.
At this point, it is unknown how HFE-7000 performs relative to other fluids in a side-
by-side comparison. Theoretically they are quite similar, but a test would need to be
49

Table 7. Comparison of Refrigerant Density


3
Fluid Density (kg/m )
State 3 State 4
100C, 600kPa 60C, 100kPa
HFE-7000 47.11 7.51
HCFC-22 17.6 3.1
HCFC-134 20.88 3.73

performed to understand how differences, including density, manifest themselves in a real-


world situation.

Cost Analysis
The estimated cost of the optimized 3kW system is approximately $11,000 (see Table
8 for breakdown). This does not include the cost of an energy source and sink, as these will
vary by application.

Table 8. System Cost Breakdown


Component Price ($)
3kW expander/generator 4500
HFE‐7000 1400
Pump 1200
Piping, valves 1000
Boiler 800
Condenser 800
Pre‐heater 600
Controller 500
Misc. hardware 500
Total 11300

For a standalone solar-powered system, a direct-heated collector (lines of black


tubing exposed to the sun) could be utilized as a source and a large shaded bladder full of
water could act as sink. To power a 3kW system at 10% efficiency, the collector field would
need to be approximately 50 square meters, assuming 800W/m2 insolation and 80%
conversion efficiency. The collector and sink could be reasonably expected to add $5,000-
$10,000 of cost, totaling roughly $20,000 for an installed 3kW system. A photovoltaic (PV)
50

system costs approximately $9W-1 installed, or $27,000 for a 3kW system (SolarBuzz, 2010),
and will require a similar area. It can be seen that a solar-powered Rankine cycle operating
at 10% efficiency is cost competitive with PV sources. However, an efficiency approaching
that level has yet to be demonstrated empirically.
For industrial waste heat applications with an existing heat source and sink the cost
increase for installation is minimal. A 3kW system that costs $12,000 operating around the
clock would have a payback period of 6.5 years at current industrial electricity prices of
$0.07 kWh-1 (Energy Information Administration, 2010).
51

CHAPTER 5

CONCLUSIONS

All primary goals of this project have been achieved:


 The use of an HFE fluid in an ORC has been demonstrated.
 The use of a low-temperature power source (97°C) has been demonstrated.
 The existing mathematical model for HFE-7000 has been both modified and validated
so that future systems can be designed using more ideal parameters such as
temperature and pressure, or other components.
 Low-cost components have been used to demonstrate potential cost effectiveness of
such a system relative to PV in solar powered applications.

SUMMARY
The use of HFE-7000 as a working fluid offers the potential to recapture energy from
a variety of existing and novel heat sources. The relatively low operating temperature of
<100°C makes it appropriate for waste heat recapture applications as well as small-scale
solar collectors. Such a system could have residential applications if a sufficiently large heat
sink were available (such as a swimming pool). The fluid is advantageous in that it is
environmentally benign and can provide useful energy at low risk and potentially low cost to
the end consumer.
Experimentation has shown that the fluid functions in a system made from off-the-
shelf components. The system achieved a thermal efficiency of 3.1% with no optimization of
components or inputs. Further theoretical analysis shows that such a system could be
optimized with alternate components to provide a net thermal efficiency as high as 10%. At
such efficiency, and without high-priced custom components, the system becomes cost
competitive with PV systems in solar powered applications, and has a short payback period
for industrial applications.
From a comparison with other commonly-used Organic working fluids, it appears that
HFE-7000 would function as a drop-in replacement fluid in other ORC systems. A similar
compressibility and heat capacity is observed in all fluids. The primary differences are a
52

higher density and vapor pressure in HFE-7000. The conclusion of interchangeability is


further strengthened by the experiment performed which required minimal retrofit to a
system that had previously run with HCFC-134.

RECOMMENDATIONS
Improvements to the present system could be achieved fairly easily:
 Optimization of the expander for various inlet pressures and flow rates could be
accomplished in a standalone experiment and greatly increase the efficiency of the
expander and overall system.
 A complete replacement of the pump is recommended. The existing pump is
inefficient even at the higher flow rates for which it was designed, causing it to
contribute thermal energy to the fluid and complicate measurements and system
analysis. A more efficient pump would mitigate these concerns.
 The addition of a second water heater would immediately increase the energy
available to the system, allowing higher flow rate and increased system performance.
Such an increase would require a commensurate increase in the heat rejection
capacity of the thermal sink, so a more massive water bath would be required.
 Increasing the temperature of the thermal sink would prevent sub-cooling in the liquid
phase and increase system efficiency.
 Improved measurement instrumentation would improve resolution of system
measurements. Thermocouple wells in the hot- and cold-water lines would allow for
instantaneous, accurate measurement of the power input to and rejection from the
system. This would eliminate the need for time-averaged bulk temperature
measurements of the source and sink, which are less accurate because the system
changes performance as the water temperatures change. An in-line flow meter to
measure flow rate of the working fluid would remove any error caused by the
calculation used to find flow rate from the energy input. More accurate flow rate
values would allow better correlation to and optimization of the expander
performance.
53

REFERENCES

3M Electronics. (2005). Novec 7000 engineered fluid product information. Retrieved April
10, 2010, from http://multimedia.3m.com/mws/mediawebserver?
mwsId=66666UuZjcFSLXTtlXftMxMVEVuQEcuZgVs6EVs6E666666.
Clemente, S., Micheli, D., Reini, D., & Taccani, R. (2010, June). Numerical model and
performance analysis of a scroll machine for ORC applications. In Proceedings of 23rd
Annual ECOS Conference. Lausanne, Switzerland.
Cogswell, F. (2006). An ORC power plant operating on a low-temperature (165 F)
geothermal source. Proceedings of the Geothermal Resources Council Annual Meeting.
Curran Associates, Inc.
Dai, Y., Wang, J., & Gao, L. (2009). Parametric optimization and comparative study of
organic Rankine cycle (ORC) for low grade waste heat recovery. Energy Conversion
and Management, 50, 576–582.
Energy Information Administration. (2010). Average retail price of electricity to ultimate
customers: Total by end-use sector. Retrieved April 10, 2010, from
http://www.eia.doe.gov/cneaf/electricity/epm/table5_3.html
Environmental Protection Agency. (2010). Emissions. Retrieved April 10, 2010, from
http://epa.gov/climatechange/index.html
Giampaoloa, M., & Leonardo, M. (2010, June). Model of a steam/organic vapour volumetric
reciprocating expander. In Proceedings of 23rd Annual ECOS Conference. Lausanne,
Switzerland.
Giampaolo, M., & Stefano, P. (2010, June). Model of a steam Wankel expander. In
Proceedings of 23rd Annual ECOS Conference. Lausanne, Switzerland.
Hettiarachchi, H., Golubovic, M., Worek, W., & Ikegami, Y. (2007). Optimum design
criteria for an organic Rankine cycle using low-temperature geothermal heat sources.
Energy, 32, 1698-1706.
Hung, T. (2001). Waste heat recovery of organic Rankine cycle using dry fluids. Energy
Conversion and Management, 42, 539-553.
Hung, T., Shai, T., & Wang, S. (1997). A review of organic Rankine cycles (ORCs) for the
recovery of low-grade waste heat. Energy, 22(7), 661-667.
Husband, W., & Beyene, A. (2008). Low-grade heat-driven Rankine cycle, a feasibility
study. International Journal of Energy Research, 32(2008), 1373-1382.
Ingley, H.A., Reed, R., & Goswami, D.Y. (2004). Optimization of a scroll expander applied
to an ammonia/water combined cycle system for hydrogen production. Retrieved April
10, 2010, from http://www.mae.ufl.edu/NasaHydrogenResearch/
pubprogramreport/Dr.%20Ingley_H2%20Production_for%20submission.pdfvv.
54

Kane, M., Favrat, D., Gay, B., & Andres, O. (2007). Scroll expander organic Rankine cycle
(ORC) efficiency boost of biogas engines. Ecos 2007, 2, 1017-1024.
Kim, H., Ahn, J., Park, I., & Rha, P. (2007). Scroll expander for power generation from a
low-grade steam source. Institute of Mechanical Engineers, 221(5), 705-711.
Lemort, V., & Quoilin, S. (2009). Designing scroll expanders for use in heat recovery
Rankine cycles. Retreived April 10, 2010, from http://www.labothap.ulg.ac.be/
cmsms/uploads/File/London_paper_proceedings.pdf.
Lemort, V., Quoilin, S., Cuevas, C., & Lebrun, J. (2009). Testing and modeling a scroll
expander integrated into an organic Rankine cycle. Applied Thermal Engineering,
29(14-15), 3094-3102.
Liu, B., Chien, K., & Wang, C. (2004). Effect of working fluids on organic Rankine cycle for
waste heat recovery. Energy, 29, 1207-1217.
Mago, P., Srinivqasa, K., Chamra, L., & Somayaji, C. (2008). An examination of exergy
destruction in organic Rankine cycles. International Journal of Energy Research, 32,
926-938.
Mago, P., Srinivqasa, K., Chamra, L., & Somayaji, C. (2008). An examination of
regenerative organic Rankine cycles using dry fluids. Applied Thermal Engineering,
28, 998-1007.
Manhire, B., (2004). On three-phase reactive power. In Proceedings of the 2004 American
Society for Engineering Educational Annual Conference and Exhibition. Salt Lake
City, Utah.
Giampaolo, M., & Stefano, P. (2010, June). Model of a steam Wankel expander. In
Proceedings of 23rd Annual ECOS Conference. Lausanne, Switzerland.
Mathias, J., Johnston, J., Cao, J., Priedeman, D., & Christensen, R. (2009). Experimental
testing of gerotor and scroll expanders used in, and energetic and exergetic modeling
of, an organic Rankine cycle. Journal of Energy Resources Technology, 131(1), 012201
(9 pages).
O’Callaghan, P., Badr, O., Probert, S. D., Bell, M. A., & Patel, R. M. (1985). Performance of
multi-vane expanders. Applied Energy, 20 (3), 207-234.
Owens, J. (1998). Segregated hydrofluoroethers: Low GWP alternatives to HFCS and PFCs.
3M Specialty Materials. Retrieved April 10, 2010, from http://multimedia.3m.com/
mws/mediawebserver?mwsId=TTTTTVv_LdGtMYu75Z0AOy9Tfwvrfdv_IwUTfwUT
fTTTTTT
Peterson, R., Wang, H., & Herron, T. (2008). Performance of a small-scale regenerative
Rankine power cycle employing a scroll expander. Proceedings of the Institute of
Mechanical Engineers, 222(3), 271-282.
Quoilin, S. (2007). Experimental study and modeling of a low temperature Rankine cycle for
small scale cogeneration. Retrieved April 10, 2010, from http://organicrankine.com/
orc_documents/scroll_non_auto/TFE_SQ010607.pdf.
55

Quoilin, S., Lemort, V., & Lebrun, J. (2010). Experimental study and modeling of an organic
Rankine cycle using scroll expander. Applied Energy, 87(2010), 1260-1268.
Saitoh, T., Yamada, N., & Wakashima, S. (2007). Solar Rankine cycle system using scroll
expander. Journal of Environmental Engineering, 2(4), 708-719.
Shigemi, N., Toshio, O., & Akira, M. (1999). Scroll expander, internal losses of expander.
Transactions of the Japan Society of Refrigerating and Air Conditioning Engineers,
10(2), 123-132.
Siddiqi, M., & Atakan, B. (2010, June). Investigation of the criteria for fluid selection in
Rankine cycles for waste heat recovery. In Proceedings of 23rd Annual ECOS
Conference. Lausanne, Switzerland.
Smith, I. Stosic, N., & Kovacevic, A. (1999). Power recovery from low cost two-phase
expanders. Centre for Positive Displacement Compressor Technology. Retrieved April
10, 2010, from http://organicrankine.com/orc_documents/screw/screw_expander.pdf.
Smith, I., Stosic, N., Kovacevic, A., & Langson, N. (2007). Cost effective small scale ORC
systems for power recovery from low enthalpy geothermal resources. Geothermal
Resources Council Transactions, 31.
Solarbuzz. (2010). Solar Module Retail Price Environment. Retrieved April 10, 2010, from
http://www.solarbuzz.com/Moduleprices.htm
Tchanche, B., Quoilin, S., Declaye, S., Papadakis, G., & Lemort, V. (2010, June). Economic
optimization of small scale organic Rankine cycles. In Proceedings of 23rd Annual
ECOS Conference. Lausanne, Switzerland.
Teagan, W., & Clay, W. (1973). 3 KW closed Rankine-cycle powerplant with a turbine
expander. Thermo Electron Corp. Retrieved April 10, 2010, from
http://www.dtic.mil/srch/doc?collection=t3&id=AD0776982
Toffoloa, A., Lazzarettoa, A., Manentea, G., & Rossi, N. (2010, June). Synthesis/design
optimization of organic Rankine cycles for low temperature geothermal sources with
the HEATSEP method. In Proceedings of 23rd Annual ECOS Conference. Lausanne,
Switzerland.
Vaja, I., & Gambarotta, A. (2010, June). Dynamic model of an organic Rankine cycle
system. Part I – Mathematical description of main components. In Proceedings of 23rd
Annual ECOS Conference. Lausanne, Switzerland.
Vaja, I., & Gambarotta, A. (2010, June). Dynamic model of an organic Rankine cycle
system. Part II – The full model: description and simulation. In Proceedings of 23rd
Annual ECOS Conference. Lausanne, Switzerland.
Wang, X., & Zhao, L. (2009). Analysis of zeotropic mixtures used in low-temperature solar
Rankine cycles for power generation. Solar Energy, 83(5), 605-613.
Yamamoto, T., Furahata, T., Arai, M., & Mori, K. (2001). Design and testing of the organic
Rankine cycle. Energy, 26, 239-251.
Yanagisawa, T., Fukuta, M., Ogo, Y., & Hikichi, T. (2001). Performance of an oil-free
scroll-type air expander. IMechE C591, (27), 167-174.
56

Zanelli, R., & Favrat, D. (1994). Experimental investigation of a hermetic scroll expander-
generator. Proceedings of the International Compressor Engineering Conference. West
Lafayette, Indiana.
Zyhowski, G., Brown, A., & Achaichia, A. (2010, June). HFC-245fa working fluid in
organic Rankine cycle - a safe and economic way to generate electricity from waste
heat. In Proceedings of 23rd Annual ECOS Conference.
57

APPENDIX

SCROLL DEVICE PROPERTIES


58

Figure 24. Compressor performance data under various conditions.


59

Figure 25. Compressor specifications.

You might also like