Sei sulla pagina 1di 135

Graduate School ETD Form 9

(Revised 12/07)

PURDUE UNIVERSITY

GRADUATE SCHOOL
Thesis/Dissertation Acceptance
This is to certify that the thesis/dissertation prepared
By Abhinav Krishna
Entitled
ORGANIC RANKINE CYCLE WITH SOLUTION CIRCUIT FOR LOW-GRADE HEAT
RECOVERY

For the degree of

Master of Science in Mechanical Engineering

Is approved by the final examining committee:


Eckhard A. Groll
Chair

Suresh V. Garimella

James E. Braun

W. Travis Horton

To the best of my knowledge and as understood by the student in the Research Integrity and
Copyright Disclaimer (Graduate School Form 20), this thesis/dissertation adheres to the provisions of
Purdue Universitys Policy on Integrity in Research and the use of copyrighted material.

Eckhard Groll
Approved by Major Professor(s): ____________________________________

____________________________________
07/25/2012

Approved by: David Anderson


Head of the Graduate Program

Date

Graduate School Form 20


(Revised 9/10)

PURDUE UNIVERSITY
GRADUATE SCHOOL

Research Integrity and Copyright Disclaimer

Title of Thesis/Dissertation:
ORGANIC RANKINE CYCLE WITH SOLUTION CIRCUIT FOR LOW-GRADE HEAT
RECOVERY

For the degree of

Master
Science
Choose of
your
degree in Mechanical Engineering

I certify that in the preparation of this thesis, I have observed the provisions of Purdue University
Executive Memorandum No. C-22, September 6, 1991, Policy on Integrity in Research.*
Further, I certify that this work is free of plagiarism and all materials appearing in this
thesis/dissertation have been properly quoted and attributed.
I certify that all copyrighted material incorporated into this thesis/dissertation is in compliance with the
United States copyright law and that I have received written permission from the copyright owners for
my use of their work, which is beyond the scope of the law. I agree to indemnify and save harmless
Purdue University from any and all claims that may be asserted or that may arise from any copyright
violation.

Abhinav Krishna

______________________________________
Printed Name and Signature of Candidate

07/12/2012

______________________________________
Date (month/day/year)

*Located at http://www.purdue.edu/policies/pages/teach_res_outreach/c_22.html

ORGANIC RANKINE CYCLE WITH SOLUTION CIRCUIT FOR LOW-GRADE


HEAT RECOVERY

A Thesis
Submitted to the Faculty
of
Purdue University
by
Abhinav Krishna

In Partial Fulfillment of the


Requirements for the Degree
of
Master of Science in Mechanical Engineering

August 2012
Purdue University
West Lafayette, Indiana

UMI Number: 1529707

All rights reserved


INFORMATION TO ALL USERS
The quality of this reproduction is dependent upon the quality of the copy submitted.
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.

UMI 1529707
Published by ProQuest LLC (2012). Copyright in the Dissertation held by the Author.
Microform Edition ProQuest LLC.
All rights reserved. This work is protected against
unauthorized copying under Title 17, United States Code

ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346

ii

ACKNOWLEDGEMENTS

The contents of this work, and my personal growth during its development, is due
in no small part to several people. My advisors Eckhard Groll and Suresh Garimella
provided plenty of guidance and support, not to mention commensurate intellectual
freedom, throughout my Masters studies. Jim Braun and Travis Horton thank you for
all your useful insights during the course of this project.
I would like to thank everyone that helped in the design and construction of the
challenging experimental setup for this project. Frank, Bob and Gilbert thank you for
your efforts. Special thanks to two tireless students, Philipp Danecker and Nick Czapla,
for the countless hours you spent working on the setup. Your motivation and competence
went a long way in raising my enthusiasm for this project.
I have had the privilege of working with some of the best graduate students
around. Brandon Woodland unassuming and soft spoken provided plenty of brain
power for this work. Not that he represents a victory of substance over style, because he
has plenty of both. Ian Bell and Craig Bradshaw provided quality mentorship and plenty
of ideas. My other colleagues at Herrick Laboratories including Bryce, Christian, Dong
Han, Howard, Huize, Simba, Stephen and several others have lent me their time, of
which they had precious little. I have been lucky to foster several friendships that extend
beyond

professional

interests

during

my

time

at

Herrick

Laboratories.

iii
Finally, I would like to thank the Cooling Technologies Research Center, and
TORAD Engineering for funding this research and providing customized equipment. I
would also like to thank ASHRAE for providing me with financial support during my
Masters studies.

iv

TABLE OF CONTENTS

Page
LIST OF TABLES ............................................................................................................. vi
LIST OF FIGURES .......................................................................................................... vii
NOMENCLATURE .......................................................................................................... ix
ABSTRACT ................................................................................................................. xii
CHAPTER 1. INTRODUCTION ......................................................................................1
1.1
Background ........................................................................................ 1
1.2
Motivation .......................................................................................... 2
1.3
Objective ............................................................................................ 3
1.4
Approach ............................................................................................ 6
CHAPTER 2. CURRENT STATUS OF TECHNOLOGY ...............................................8
2.1
Organic Rankine Cycles .................................................................... 8
2.2
Absorption Power Cycles .................................................................. 9
2.3
Working Fluid Mixtures .................................................................. 11
CHAPTER 3. THERMODYNAMIC MODEL DEVELOPMENT
AND RESULTS .......................................................................................13
3.1
Baseline Cycles ................................................................................ 13
3.1.1
Organic Rankine Cycle ............................................................ 13
3.1.2
Vapor Compression Cycle with Solution Circuit ..................... 14
3.1.3
Organic Rankine Cycle with Solution Circuit .......................... 16
3.2
Thermodynamic Features of Binary Mixtures ................................. 18
3.2.1
Phase Equilibrium .................................................................... 18
3.2.2
Absorption / Desorption Process .............................................. 20
3.2.3
Temperature Glide and Capacity Control ................................ 22
3.3
Cycle Model of an Organic Rankine Cycle with Solution Circuit .. 25
3.3.1
Mass Balance ............................................................................ 27
3.3.2
Energy Balance......................................................................... 28
3.4
Organic Rankine Cycle with Solution Circuit Model Results ......... 32
CHAPTER 4. DESIGN OF EXPERIMENTAL TEST SYSTEM ..................................45
4.1
System Sizing................................................................................... 45
4.2
System Design and Layout .............................................................. 48
4.3
Design and Selection of Major Components ................................... 53
4.3.1
Pump ......................................................................................... 53
4.3.2
Heat Exchangers ....................................................................... 57
4.3.3
Expander ................................................................................... 60
4.3.4
Separator ................................................................................... 64

v
Page
4.3.5
4.3.6
4.3.7

Receiver .................................................................................... 66
Expansion Valves ..................................................................... 66
Instrumentation and Data Acquisition ...................................... 67
4.4
Experimental Error and Uncertainty ................................................ 68
CHAPTER 5. EXPERIMENTAL RESULTS AND ANALYSIS...................................71
5.1
Experimental Program Overview .................................................... 71
5.2
Experimental Performance Trends .................................................. 75
5.3
Experimental Program Summary..................................................... 81
CHAPTER 6. PERFORMANCE CHARACTERIZATION OF THE
ORCSC SYSTEM ....................................................................................83
6.1
Parametric Analysis ......................................................................... 83
CHAPTER 7. CONCLUSIONS AND FUTURE WORK...............................................90
7.1
Conclusion ....................................................................................... 90
7.2
Recommendations for Future Work................................................. 92
LIST OF REFERENCES ...................................................................................................95
APPENDICES
Appendix A Instrument Calibration Procedures .................................................. 98
Appendix B
Wiring Diagram for Data Acquisition System............................... 105
Appendix C
Procedures for Charging and Discharging the ORCSC System .... 106
Appendix D Procedures for Operating the ORCSC System .............................. 108
Appendix E
Data from Experimental Testing .................................................... 112

vi

LIST OF TABLES

Table ..............................................................................................................................Page
Table 3-1: Key assumptions used in the thermodynamic model. ..................................... 32
Table 4-1: Sample set of input parameters used as a basis for system design. ................. 46
Table 4-2: Sample property values at each state point in the
ORCSC cycle using the LKP EOS (refer to Figure 3-3). ................................46
Table 4-3: Capacities, flow rates and cycle efficiencies used as a design
basis for the ORCSC. .......................................................................................47
Table 4-4: Qualitative comparison for different pump types for an ORCSC
application. .......................................................................................................54
Table 4-5: Primary specifications of CAT PUMP MODEL 1051.CO2. .......................... 56
Table 4-6: Data acquisition devices and their output signals. .......................................... 68
Table 4-7: Measurement uncertainties. ............................................................................. 70
Table 5-1: Experimental test matrix. ................................................................................ 72
Table 5-2: Key assumptions used in the thermodynamic model for
experimental comparison. ................................................................................75
Table 6-1: Baseline conditions for parametric analysis. ................................................... 84
Table E-1: Experimental temperature data. .................................................................... 112
Table E-2: Experimental temperature data continued. ................................................... 113
Table E-3: Experimental pressure data. .......................................................................... 114
Table E-4: Experimental flow rate data. ......................................................................... 115
Table E-5: Experimental power, performance and heat transfer data. ........................... 116

vii

LIST OF FIGURES

Figure .............................................................................................................................Page
Figure 1-1: Heat sources, their temperature levels and heat recovery technologies. .......... 4
Figure 3-1: Schematic representation of the Organic Rankine Cycle............................... 13
Figure 3-2: Schematic representation of the Vapor Compression Cycle
with Solution Circuit. .....................................................................................15
Figure 3-3: Schematic representation of the Organic Rankine Cycle
with Solution Circuit. .....................................................................................17
Figure 3-4: p, T, diagram for an arbitrary mixture (modified from Kyle 1999). ........... 20
Figure 3-5: Variation of liquid and vapor composition with temperature
for a binary mixture. .......................................................................................21
Figure 3-6: Irreversibilities in heat transfer processes (modified from Mulroy 1993). .... 24
Figure 3-7: Impact of ORCSC concentration pairings on efficiency for a
source temperature of 100C and a sink temperature of 20C. ......................33
Figure 3-8: Variation of overall Second Law Efficiency as a function of
Circulation Ratio. ...........................................................................................34
Figure 3-9: Impact of rich solution concentration on efficiency for various source
temperatures. Plot is optimized for weak solution concentration. .................35
Figure 3-10: Second Law efficiency as a function of source temperature
for a basic ORC.............................................................................................36
Figure 3-11: Second Law efficiency as a function of source temperature
for an ORC with internal regeneration at the expander outlet. .....................36
Figure 3-12: Second Law efficiency as a function of source temperature
for the ORCSC. .............................................................................................37
Figure 3-12: Variation of component capacities as a function of Circulation Ratio. ....... 40
Figure 3-13: Variation of High-side system Pressure as a function of
Circulation Ratio. ..........................................................................................41
Figure 3-14: Variation of Pressure Ratio as a function of Circulation Ratio.................... 41
Figure 4-1: Detailed schematic diagram of the secondary loop for heat input
temperature control.........................................................................................49
Figure 4-2: Detailed schematic diagram of the primary ORCSC system. ........................ 49
Figure 4-3: Preliminary CAD representation of ORCSC system arrangement. ............... 51
Figure 4-4: Top view of the primary ORCSC experimental load stand. .......................... 52
Figure 4-5: Front view of the primary ORCSC experimental load stand. ........................ 52
Figure 4-6: Secondary heat input loop load stand. ........................................................... 53
Figure 4-6: Sectional view of CAT PUMP MODEL 1051.CO2. ..................................... 56
Figure 4-7: Dual Coil Heat Exchangers selected for use in the ORCSC. ......................... 58

viii
Figure .............................................................................................................................Page
Figure 4-8: Sanden TRS-105 scroll compressor housing. ................................................ 62
Figure 4-9: Front view of Torad rotary spool expander housing. ..................................... 63
Figure 4-10: Design of the Torad rotary spool expander. ................................................. 64
Figure 4-11: Design and layout of separator and sight glass. ........................................... 65
Figure 4-12: Receiver and sight glass selected for use in the ORCSC. ............................ 66
Figure 4-13: General overview of the data aquistion system. .......................................... 67
Figure 5-1: Comparison of theoretical and experimental Second Law
efficiencies as a function of the Circulation Ratio. ........................................77
Figure 5-2: Comparison of theoretical and experimental Desorber Heat
Input (System Capacity) as a function of the Circulation Ratio.....................78
Figure 5-3: Comparison of theoretical and experimental Second Law Efficiencies
as a function of the Desorber Exit Temperature. ............................................78
Figure 5-4: Comparison of theoretical and experimental Expander Pressure Ratio
as a function of the Circulation Ratio.............................................................79
Figure 5-5: Comparison of theoretical and experimental Pump Work Input as a
function of the Circulation Ratio. ...................................................................80
Figure 5-6: Comparison of theoretical and experimental Expander Work Output as a
function of the Circulation Ratio. .................................................................. 80
Figure 6-1: Net power output and Second Law Efficiency as a
function of expander isentropic efficiency. ....................................................85
Figure 6-2: Net power output and Second Law Efficiency as a function of pump
efficiency. .......................................................................................................86
Figure 6-3: Net power output and Second Law Efficiency as a function of the
desorber pinch point temperature. ..................................................................87
Figure 6-4: Net power output and Second Law Efficiency as a function of the
absorber pinch point temperature. ..................................................................88
Figure 6-5: Net power output and Second Law Efficiency as a function of the
internal heat exchanger effectiveness. ............................................................89
Figure A 1: Sample calibration data for a thermocouple. ................................................. 99
Figure A 2: Sample calibration data for a pressure transducer. ...................................... 101
Figure A 3: Sample calibration data for a mass flow meter. .......................................... 102
Figure A 4: Top view of expander and torque cell. ........................................................ 103
Figure A 5: Calibration of the torque cell. ...................................................................... 104
Figure A 6: Sample calibration data for torque cell. ....................................................... 104

ix

NOMENCLATURE

Symbols

area

CR

Circulation Ratio,

degrees of freedom

gravitational acceleration

enthalpy

IHX

Internal Heat Exchanger

number of components

mg

mass flow rate of glycol

mr

mass flow rate of rich solution

mw

mass flow rate of weak solution

mv

mass flow rate of vapor

pressure

capacity

QA

absorber capacity

QD

desorber capacity

Qh , w

heating capacity for water

Qh ,tot

total heating capacity

temperature

volumetric flow rate

x
w

overall uncertainty

work

quality

measured quantity

difference

heat exchanger effectiveness

efficiency

relative humidity

density

number of phases

concentration

specific volume

Tor

torque

rotational speed

Subscripts
a

air

absorber

cold fluid

car

carnot

comp

compressor

cond

condensing

crit

critical

desorber

hot fluid

inlet

IHX

Internal Heat Exchanger

meas

measured

outlet

xi
pump

pump

pinch

pinch-point temperature

rich solution

ref

refrigerant

isentropic

sat

saturation

sub

sub-cooling

sup

superheat

turb

turbine

tot

total

weak solution

vapor

Acronyms
DAQ

Data Aquisition

EOS

equation of state

EPDM

Ethylene Propylene Diene Monomer

LKP

Lee-Kesler-Plcker equation of state

ORC

Organic Rankine Cycle

ORCSC

Organic Rankine Cycle with Solution Circuit

PR

Pressure Relief

PS

Pressure Switch

PTFE

Polytetrafluoroethylene

VCCSC

Vapor Compression Cycle with Solution Circuit

VLE

vapor-liquid-equilibrium

xii

ABSTRACT

Krishna, Abhinav. M.S.M.E., Purdue University, August 2012. Organic Rankine Cycle
with Solution Circuit for Low-Grade Heat Recovery. Major Professors: Eckhard A. Groll,
Suresh V. Garimella.

Increasing interest in utilizing low-grade heat for power generation has prompted
a search for ways in which the power conversion process may be enhanced. A novel
Organic Rankine Cycle with Solution Circuit (ORCSC) using Carbon Dioxide / Acetone
as the working fluid pair was studied for this purpose. A thermodynamic simulation
model was developed and an experimental test stand was built to serve as a proof of
concept for the technology.
The thermodynamic model showed that the ORCSC using Carbon Dioxide /
Acetone as the working pair offers no significant efficiency improvements over a
conventional Organic Rankine Cycle (ORC) using only Carbon Dioxide as the working
fluid. Furthermore, the ORCSC with a Carbon Dioxide / Acetone working pair has
significantly lower performance than an ORC using conventional working fluids such as
pentane or R245fa. This may render the ORCSC unattractive since the low-temperature
heat sources mean that the theoretical (Carnot) efficiency limit is itself relatively low, and
achieving cycle efficiencies as close to the Carnot limit as possible is necessary for
ensuring the economic feasibility of the technology. However, the ORCSC was

xiii
found to have significantly lower working pressures than an ORC, provides the ability to
use temperature glide to match the temperature profiles of the source and sink fluids and
facilitates intrinsic capacity control. This may lead to higher overall system efficiencies
when coupled with sources that have varying heat input temperatures or loads. More
application-specific studies that address the nature and capacity of the source and sink
streams are required to identify where this ability may be most advantageous.
The experimental tests showed good agreement with the simulation data when all
the boundary conditions were matched. However, the efficiencies of the system were
generally poor and many of the expected trends were skewed due to design shortcomings
and the use of equipment that was not optimized for the ORCSC system. Isolation of
individual parameters was an acute challenge due to the number of variables that need to
be tightly controlled during system operation. Nevertheless, the experimental results
provided a validation of the simulation model.
The simulation model was expanded to include a parametric study of the various
components on the overall system performance. It showed that the ORCSC is particularly
sensitive to the performance of the expander and pump. Amongst the heat exchangers,
the performance of the absorber had the greatest impact on the overall system
performance.
It is clear from this study that a range of practical considerations need to be taken
into account and weighed together with the thermodynamic analysis when evaluating the
feasibility of ORCSC technology. The ORCSC offers some potential practical advantages
which may outweigh the added cost and complexity of these systems in certain
applications. However, the maturity of the technology and associated body of literature is

xiv
limited, and further work needs to be pursued in this area before widespread adoption of
the technology is possible.

CHAPTER 1. INTRODUCTION

1.1 Background
The increasing cost of energy, coupled with the recent drive for energy security
and climate change mitigation have provided the impetus for harnessing renewable
energy sources as viable alternatives to conventional fossil fuels. However, several of
these renewable energy sources, including geothermal, biomass and solar, intrinsically
provide low-grade heat (at temperatures between 60C 300C). Furthermore,
thermodynamic considerations mean that a large amount of low-grade heat is discharged
from power plants and various other industrial processes. In fact, in the United States,
over two-thirds of the primary energy supply is ultimately rejected as low-grade waste
heat according to the World Energy Council (2006). Recovering low-grade heat,
therefore, is increasingly becoming an economic and environmental imperative.
Low-grade waste heat, largely at a temperature level between 30C 250C, is
primarily a product of thermo-mechanical energy conversion losses. Thus far, waste heat
recovery systems have mainly been designed for thermal heat recovery for process use
using recuperative heat exchangers. Conversions from waste heat to higher forms of
energy, such as shaft work and/or electricity, have not been viable due to technical and
cost impediments. Nevertheless, due to increasing energy costs and available heat sources,

2
waste heat recovery systems are of increasing interest to designers, engineers and society
at large (Little 2009).
1.2 Motivation
An increasing source of waste heat comes from growing technology needs. In the
last twenty years, computer power consumption has increased exponentially. Of the total
electricity consumption in the United States in 2006, more than 1 % was used to operate
large data centers (Koomey 2007). According to ASHRAE (2005), datacom workload is
expected to rise further at a 40 to 50 % compound growth rate. Additionally, the power
density of datacom equipment is expected to reach up to 8 kW/m2 for computer servers
and 15 kW/m2 for high density communication equipment by 2014. In large data centers,
therefore, an enormous amount of electrical energy is consumed, which is directly
converted into heat and ultimately rejected to the environment.
The rejected heat represents a large energy stream that has already been paid for;
however recovering and utilizing it represents a challenge. This is primarily due to the
temperature level of the waste heat (50C 85C), which renders it inefficient to convert
to a more usable form of an energy transport medium, such as electricity, using
conventional power generation technology. Furthermore, several renewable energy
sources face the same problem low source temperatures to the point where the
application of steam Rankine Cycles, ubiquitous in power generation, is grossly
inefficient and expensive.
Currently, technologies attempting to provide low-grade heat recovery solutions
have seen very limited commercialization. This is broadly due to two reasons: lack of

3
historical research and development in the area of waste heat recovery due to technical
and cost impediments; and technical challenges associated with scaling the technology
from utility scale to commercial scale, particularly with regard to expansion machines
(turbines). However, due to rising primary energy costs and the environmental premium
being placed on fossil fuels, the conversion from low-grade heat to electrical energy is a
pressing societal challenge.

1.3 Objective
One way to recover low-grade heat, and use it in a power generating capacity, is
to use Organic Rankine Cycle (ORC) technology. Organic Rankine Cycles differ from
traditional steam Rankine Cycles in the use of an organic working fluid as opposed to
water/steam as the working fluid, making them far better suited for low temperature heat
sources. However, due to the low-temperature heat sources, the theoretical (Carnot)
efficiency limit is relatively low. Therefore, achieving cycle efficiencies as close to the
Carnot limit as possible remains a challenge, and is important for ensuring the economic
feasibility of the technology.

An overview of available heat sources and their

temperature levels, as well as the technologies that could be used for a given temperature
range to recover the available heat, is given in Figure 1-1.

Figure 1-1: Heat sources, their temperature levels and heat recovery technologies.
In order to achieve cycle efficiencies as close to the Carnot limit as possible, it
may be necessary to modify the Organic Rankine Cycle. One novel modification,
proposed by Maloney and Robertson (1953) as well as Kalina (1983), is to introduce
absorption technology for power generation. This modification is termed the Organic
Rankine Cycle with Solution Circuit (ORCSC), also known as the Absorption-Rankine
cycle. The ORCSC differs from the ORC primarily through the use of a zeotropic
mixture, consisting of a refrigerant and an absorbent as the working fluids. The
refrigerant and absorbent are characterized by a large boiling point difference. This
enables the separation of the more volatile component (the refrigerant) in the vapor phase

5
from the absorbent solution in the liquid phase. The refrigerant vapor (and a small
quantity of the solution) then flows through the expansion device, whereas the liquid
absorbent forms a solution circuit.
Aside from the possibility of higher exergetic conversion efficiency than an ORC,
the ORCSC has several other inherent features that address critical issues related to the
general applicability of low-grade heat recovery technology. For example, the solution in
the ORCSC ensures significantly lower working pressures when compared to a
conventional ORC. This allows for the use of high-pressure natural refrigerants, such as
carbon dioxide, at moderate operating pressures which leads to significant cost savings.
The use of natural refrigerants has numerous environmental benefits, including negligible
Global Warming Potential (GWP), zero Ozone Depletion Potential (ODP), and nontoxicity. The use of a zeotropic mixture in the ORCSC also means that the system
capacity can be easily adjusted by simply changing the concentration of the refrigerant
used in the system. This offers a simple, cost-effective solution for adapting the system
for peak and non-peak loads, which the system is bound to encounter in practice.
Furthermore, as shown in Figure 1-1, the added system complexity of the ORCSC can be
justified at two ends of the source temperature spectrum: at extremely low source
temperatures (<100C), where achieving cycle efficiencies as close to the Carnot limit as
possible is important for ensuring the economic feasibility of the technology; and at
intermediate source temperatures, between 200C 400C, where the source temperature
is too low for steam based cycles to be efficient, and too high to the point where
traditional lubricants used in ORC expansion machines start to break down. The solution
carryover in the ORCSC can act as the lubricant in this case. The intermediate source

6
temperature range is important because it fills an important technology gap where
existing conversion systems cannot efficiently generate power. Numerous exhaust heat
streams from industrial processes, as well as biomass and certain solar concentration
techniques, fall into this spectrum.
Despite significant advantages, there have been few ORCSC prototypes that have
been built. The few prototypes that exist have relied on using an Ammonia / Water
working pair, which renders them unattractive in many applications due to the corrosive
properties and toxicity of Ammonia. In this work, the objective is to investigate an
ORCSC cycle that uses a natural refrigerant in the working pair, and to demonstrate a
proof-of-concept experimental system.

1.4 Approach
In order to fulfill the objectives of this project, the following approach was
adopted in the given order:

Literature and patent search of Organic Rankine Cycle technology.

Identification of methods and measures from the literature to increase the


efficiency and applicability of Organic Rankine Cycles, including novel cycle
modifications.

Evaluation of different working fluids.

Thermodynamic simulations of possible cycles and assessment of their feasibility.

Combination of the identified ideas for Organic Rankine Cycle modifications,


including an Organic Rankine Cycle with Solution Circuit.

Selection of the most promising working pair.

Detailed simulations and selection of the most promising concept.

Design and construction of an experimental bread board system based on


simulation results to serve as a proof of concept for the technology.

Implementation of the measurement system and development of the test software.

Experimental tests according to a predetermined test matrix, and determination of


several key system characteristics such as efficiency, capacity, etc.

Comparison of measured performance to simulation results.

Model refinements based on test results.

Feasibility assessment of the technology, and recommendations for improvements


to the system based on experimental experience.

Identification of future technology development needs.

CHAPTER 2. CURRENT STATUS OF TECHNOLOGY

2.1

Organic Rankine Cycles

Organic Rankine Cycles with low temperature heat input are relatively well
known and have been investigated widely in the literature. Theoretical investigations
were pursued as early as the 1970s by Davidson (1977) for integration with solar
collectors, and further expanded by Probert et al. (1983). Experimental investigations
were conducted by Monahan (1976), with reported First Law thermal efficiencies usually
below 10% for small-scale systems. The experimental investigations identified that
expansion turbines suitable for use in ORC systems have not been widely studied, and
few commercial designs were available. In general, experimental investigations have
largely involved the use of vane expanders (Badr et al. 1990, Davidson 1977).
Hung et al. (1997) compared the efficiencies for various ORC working fluids such
as benzene, ammonia, R11, R12, R134a and R113. The study established correlations
between system efficiencies, source temperatures and system pressures. Of the fluids
investigated, benzene was found to provide the highest efficiency, followed by R113,
R11, R12, R134a, and Ammonia.

9
Despite the high Ozone Depletion Potential (ODP) and Global Warming Potential
(GWP) of many of these fluids, the first commercial applications appeared in the late 70s
and 80s with medium-scale power plants developed for geothermal and solar applications.
Currently, over 300 ORC systems are in operation worldwide, with over 1800 MWe of
installed capacity (and this number continues to grow at an ever increasing pace). The
largest number of plants is installed for biomass Combined Heat and Power (CHP)
applications, followed by geothermal plants and then Waste Heat Recovery (WHR)
plants (Quoilin 2011). It should be noted, however, that the largest application in terms of
installed power are geothermal applications. (Enertime 2011).

2.2

Absorption Power Cycles

The foundation for the Organic Rankine Cycle with Solution Circuit (ORCSC),
also known as the Absorption-Rankine Cycle, is the Vapor Compression Cycle with
Solution Circuit (VCCSC), first investigated by Altenkirch (1950). Groll and
Radermacher (1994) carried out successful experimental testing to demonstrate the
concept. Reversing the VCCSC creates a power generating cycle similar to the Rankine
Cycle, but with higher potential efficiencies (Kalina 1983). While Absorption-Rankine
Cycles have been known for more than 50 years, limited research and even more limited
experimental investigations have been carried out in this field.
One experimental investigation was performed by Maloney and Robertson (1953)
using an Ammonia / Water pair as the working fluid. Their results showed that the
absorption power cycle had no thermodynamic advantage over the Rankine Cycle.

10
However, the authors encouraged further investigations in this field, and proposed the use
of other binary mixtures.
Further investigations were conducted by Kalina (1983) with the same working
fluid as Maloney and Robertson, but with a slightly different experimental setup. Kalina
showed that the cycle has a thermal efficiency that is 30-60% greater than comparable
steam power cycles at the same source temperature. In these studies, the cycle was
coupled with relatively high turbine inlet temperatures of 180C.
Due to the diversity of potential applications, the Kalina cycle was further studied
for different purposes with slightly different configurations. Goswami and Xu (2000)
proposed a simple combined cycle using solar energy as the heat source. Zheng et al.
(2006) modified the Kalina cycle in order to produce power as well as provide
refrigeration simultaneously, but the investigations carried out were numerical in nature,
without experimental validation.
Robbins and Garimella (2010) published theoretical investigations of an Organic
Rankine Cycle with Solution Circuit using a novel binary mixture of Amyl-Acetate and
Carbon Dioxide (CO2) as the working fluid. Their parametric theoretical results showed
promising thermal efficiencies, but once again, experimental validation was not carried
out.
In summary, while there have been a few experimental investigations conducted
for Absorption-Rankine Cycles using Ammonia / Water as the working pair, to the best
of the authors knowledge, no investigations to date have considered novel working pairs
in an experimental investigation of the Absorption-Rankine Cycle.

11
2.3

Working Fluid Mixtures

The choice of a working fluid mixture for the ORCSC is defined by the
requirement that the two fluids have a large boiling point difference. This enables the
more volatile component (refrigerant) to easily separate from the liquid absorbent, and
for the system to accommodate a large range of source and sink temperatures simply by
adjusting the composition of the mixture. Generally, the working fluid mixtures found in
absorption cycles may also be used in the ORCSC. Ammonia / Water mixtures have been
studied extensively in the literature; however, ammonia has the obvious drawback of
being toxic and corrosive, which limits its potential applications. An Amyl-Acetate /
Carbon Dioxide mixture was studied for use in an ORCSC (Robbins and Garimella,
2010); however, the operating pressures were found to be too high for this working pair.
Groll and Radermacher (1994) studied the use of a Carbon Dioxide / Acetone mixture in
a VCCSC application, and Carbon Dioxide was found to have the following advantages:

Low Global Warming Potential and zero Ozone Depletion Potential as compared
to conventional refrigerants. This is pertinent given that the application of this
technology ultimately focuses on mitigating environmental impact.

Non toxicity.

Non-flammability in this working pair, CO2 can be considered an ideal fire


extinguishing medium for the Acetone. In case of a leak, a large amount of CO2
and a small quantity of Acetone would escape from the system due to the
difference in the vapor pressures of the two components.

Large volumetric heat capacity, which enables the use of smaller turbines and
other components.

12

Compatibility with common component materials.

Simplicity of operation.

No recycling of working fluid required.

Furthermore, Acetone was chosen as the absorbent solution for the following reasons:

Higher overall thermodynamic efficiencies due to its ability to dissolve CO2.

Wide availability at low cost.

Lower flammability when compared to other hydrocarbon solutions.

Given these advantages, a compelling case can be made for a Carbon Dioxide / Acetone
working pair to be investigated in an ORCSC system.

13

CHAPTER 3. THERMODYNAMIC MODEL DEVELOPMENT AND RESULTS

3.1
3.1.1

Baseline Cycles
Organic Rankine Cycle

The conceptual foundation for the ORCSC lies in a combination of a conventional


Organic Rankine Cycle with a Vapor Compression Cycle with Solution Circuit (VCCSC).
A simplified schematic diagram of an ORC is shown in Figure 3-1 below.

Low-Grade Heat Source TH

Evaporator

(1)

Expander
(4)
Condenser

Pump

(2)

(3)

Environment TL

Figure 3-1: Schematic representation of the Organic Rankine Cycle.


The ORC is essentially made up of the four main components found in traditional
steam Rankine Cycles: an evaporator, expander, condenser, and pump. However, unlike
in steam Rankine cycles, there is usually no water-steam separation drum connected to

14
the boiler, and one single heat exchanger is used to perform all three evaporation phases:
preheating, vaporization and superheating. Due to a combination of cost impediments,
system scale and thermodynamic properties of the chosen working fluid, reheating and
turbine bleeding are generally not suitable for the ORC. However, an internal regenerator
installed at the expander outlet is often used to preheat the liquid from the pump outlet.
The working principle is as follows: the refrigerant leaves the evaporator as a
supercritical gas (1), which then enters an expander. The thermal expansion through the
expander produces mechanical shaft power (2), which can be converted to electrical
energy in a generator. The refrigerant then enters the condenser, where it is converted to
the liquid phase by rejecting heat to the ambient (3). The liquid refrigerant is then
pumped back to the evaporator inlet to complete the cycle (4).

3.1.2 Vapor Compression Cycle with Solution Circuit


The VCCSC combines absorption and compression technology, and differs from the
conventional vapor compression cycle primarily by employing a working fluid mixture
consisting of a refrigerant and an absorbent, instead of pure components. A schematic
diagram of the VCCSC is provided in Figure 3-2.

15
Heat Sink TL
Rich Solution

(4)
(7)

Expansion
Valve
(5)

Solution
Pump

(3) Absorber
Internal Heat
Exchanger

(2)

(8)
Compressor
(1)

Weak
Solution

Desorber
(6)

Heat Source TH

Figure 3-2: Schematic representation of the Vapor Compression Cycle with Solution
Circuit.
The mixture is evaporated in the desorber using a heat source (which may be the
conditioned space); however, the evaporation of the mixture is not complete. Instead, a
liquid and vapor exist at the same temperature and pressure (with differing concentrations)
in the desorber. The refrigerant-rich vapor and low-concentration liquid (i.e. weak
solution) are separated at the desorber outlet. Note that an additional separator may be
necessary to complete the separation process. While the vapor (1) proceeds to the
mechanically driven compressor, the weak solution liquid (6) is pumped to a recuperative
internal heat exchanger (7) where it is used to preheat the rich solution from the absorber
(4). At the other end of the cycle, the compressed refrigerant (2) and weak solution (8)
enter the absorber. Since absorption is an exothermic process, a heat and mass exchange
process takes place in which heat is rejected to the environment (or other appropriate heat
sink) and the refrigerant is simultaneously resorbed into the absorbent, thereby forming a

16
rich solution (3). An expansion valve is used to equalize the high-side and low-side
pressures (5).

3.1.3

Organic Rankine Cycle with Solution Circuit

The ORCSC reverses the VCCSC and applies the operating principles of an ORC
to create a power generating cycle based on the use of a binary mixture. Figure 3-3 shows
a schematic diagram of the Organic Rankine Cycle with Solution Circuit. State (1)
represents the outlet of the desorber (note that a separator may be used to separate the
vapor and liquid streams at the desorber outlet). At this state, the heat source has heated
the mixture, and the primary working fluid is desorbed from the solution. The CO2 vapor
stream then enters the expander, where it is expanded to its low pressure state (2) while
producing mechanical shaft power. State (3) represents the outlet of the absorber, where
the CO2 has been resorbed into the solution to form a rich solution. Since absorption is an
exothermic process, the absorber rejects heat to the environment during this process.
Following this, the rich solution is pumped to the high pressure state (4) by means of a
solution pump, and is subsequently preheated by an internal heat exchanger (5) before
entering the desorber. State (6) represents the liquid phase weak solution at the desorber
outlet. The weak solution stream is then subcooled (7) through the internal heat
exchanger, and expanded to the low pressure state (8) by an expansion valve.

17
Low-Grade Heat Source TH
Rich Solution

(1)

Desorber

(5)

(6)
Internal Heat
Exchanger

(4)

Weak
Solution

Expander

(7)

Pump

(2)

Absorber
(3)

(8)

Expansion
Valve

Environment TL

Figure 3-3: Schematic representation of the Organic Rankine Cycle with Solution Circuit.
The main difference between the conventional vapor compression cycle and ORC
when compared to the VCCSC and ORCSC, respectively, is the use of a zeotropic
mixture with a large boiling point difference instead of a pure fluid. By introducing such
a working fluid mixture, three important features are accomplished:
1) Although desorption and absorption occur at constant pressures, the saturation
temperatures are no longer constant but vary with the composition changes of the
liquid and the vapor phases which occur during the phase change processes. This
results in a temperature glide in the absorber and desorber. These temperature
glides can be adjusted over a wide range or eliminated almost entirely.
2) A change in the overall concentration of the mixture circulating through the cycle
results in a change of the vapor pressures and densities at a given temperature,
and therefore in a change of the capacity of the entire unit.

18
3) By introducing a solution in the cycle, the operating pressures of the cycle are
significantly reduced. The solution allows the resorption temperature of the
mixture to be higher than the critical temperature of the pure refrigerant.
All of these features can be used to increase the overall COP and expand the
flexibility of system operation. However, in order to understand these features more fully
and evaluate their applicability in a power generating cycle, a detailed study of
multiphase-multicomponent systems is necessary.

3.2

Thermodynamic Features of Binary Mixtures

This section provides a background on the thermodynamic treatment of binary


mixtures. Since the proposed cycle utilizes a two component mixture, emphasis is placed
on binary working fluids; however, the same concepts would hold if multicomponent
mixtures were considered. A basic understanding of binary mixture behavior is essential
to understand features such as temperature glide and capacity control that are integral
facets of the ORCSC.

3.2.1

Phase Equilibrium

A single fluid is considered to be in phase equilibrium when successive pressure


and temperature measurements of the liquid and vapor phase do not vary with time. For a
binary mixture to be in phase equilibrium, in addition to pressure and temperature, the
concentration of each component may not vary with time. In general, the Gibbs phase
rule ( 3-1 ) gives the degrees of freedom of a system:

19

F = K + 2

( 3-1 )

The degrees of freedom (F) depends on the number of components (K), and on the
number of phases ( ) that are prevalent. When F number of intensive variables are
specified, the system is determinate. Therefore, in order to visualize the phase behavior of
a binary system, three intensive variables need to be considered when considering a
single-phase region (two intensive variables are needed when considering the two-phase
region). Typically, the variables chosen are the temperature (T), pressure (p) and the mass
concentration ( ). In the context of this work, when considering a binary mixture
consisting of a primary working fluid and an absorbent, the mass concentration may be
defined as:

liquid
=

massofprimary working fluid in theliquidphase

massofbothcomponetsintheliquidphase

vapor
=

massofprimaryworking fluid inthevaporphase

massofbothcomponetsinthevaporphase

( 3-2 )

( 3-3 )

For the mass concentration of a binary mixture, the knowledge of either the vapor
or liquid phase is sufficient, because the corresponding mass fractions can be derived
from the knowledge of the other. Figure 3-4 represents a p, T, diagram for an arbitrary
mixture.

20

Figure 3-4: p, T, diagram for an arbitrary mixture (modified from Kyle 1999).
The thick, solid line represents the saturated liquid boundary; any point above the
line would indicate that the binary mixture is in the subcooled region and a single liquid
phase exists. Similarly, the dashed line corresponds to the saturated vapor boundary; any
point below the line would indicate that the binary mixture is in the superheated phase.
For a point located in between the two boundaries, an equilibrium between the liquid and
vapor phases exists. Therefore, the pressure and temperature are the same for both phases,
while concentrations differ in each phase. This is discussed in further detail below.

3.2.2

Absorption / Desorption Process

In order to understand the thermodynamic behavior of binary mixtures more fully,


it is necessary to illustrate the absorption and desorption processes in detail. Figure 3-5

21
shows a two-dimensional rendition of Figure 3-4 where the pressure is held constant, i.e.,
the horizontal plane in Figure 3-4.

Figure 3-5: Variation of liquid and vapor composition with temperature for a binary
mixture.
Point 1 indicates a subcooled binary mixture in the liquid phase for a given mass
concentration (1), temperature (T1) and pressure (p). As the solution is heated at constant
pressure, point 1 moves up vertically until it reaches the saturated liquid line. The first
vapor bubbles begin to form at temperature T2. At equilibrium, a horizontal line ties
together the saturated liquid and vapor curves. These lines, called tie lines, reflect the

22
equilibrium between liquid and vapor compositions (, and , ) at a constant pressure

and temperature for a binary mixture. At point 2, the mass fraction of the liquid phase is
exactly the same as for the subcooled liquid at point 1(l). The mass fraction of the vapor
phase is given by v,2, where the pressure and temperature are the same as the liquid
phase as a condition for equilibrium. Note that the vapor concentration is rich in the
primary fluid, because the boiling temperature ( Tboil,primary fluid ) of the pure refrigerant is
lower than the boiling point of the pure absorbent ( Tboil,absorbent ). As the mixture is heated
further at constant pressure to state point 3, the amount of primary fluid (l,3) remaining
in liquid phase is less than at state point 2 (l). Accordingly, the vapor phase (v,3) also
contains a higher mass fraction of absorbent and a lower mass concentration of the
primary fluid compared to state point (2). Continued heating eventually moves the
mixture to point 4, where the last remaining droplet is evaporated. At this point, the vapor
has exactly the same mass concentration (v,4) as the sub cooled liquid (l). Further
heating leads to a superheated vapor (5) at the same concentration.
To have a better analytical understanding of two-phase behavior, the phase rule
( 3-1 ) can be applied. Given a binary mixture (K=2) and a two-phase region ( = 2) at
constant pressure, only one degree of freedom remains. Therefore, either the temperature
(T) or the mass concentration () can be selected to fix the equilibrium state.

3.2.3

Temperature Glide and Capacity Control

A key feature of utilizing a mixture-based cycle is that although desorption and


absorption occur at constant pressures, the saturation temperatures are no longer constant

23
but vary with the composition changes of the liquid and the vapor phases that occur during
the phase change processes. For example, during the evaporation process illustrated in
Figure 3-5, the saturation temperature changes from T2 to T4. This temperature glide
becomes larger as the difference in the boiling points of the pure components increases.
The temperature glide also depends on the mass fraction and the shape of the vapor
bubble, which is an intrinsic property of the working fluid mixture. In general, the
temperature glide is larger for intermediate values of initial than for small or large
values of . An important feature is that the temperature glide can be adjusted over a wide
range; for example, for a value of 1 in Figure 3-5, the first bubbles form at a higher
saturation temperature (2 ) when compared to the initial case of T2 and 1. Additionally,

the evaporation process is completed at a higher temperature (4 ). As a result, the


saturation temperatures are shifted to higher temperature levels with a smaller

temperature glide simply by varying the mass concentration. The same effect can be
accomplished by having an incomplete evaporation process in which a liquid and vapor
exist at the same temperature and pressure, but at different concentrations.
A key feature of the temperature glide is that it can reduce heat transfer
irreversibilities in the heat exchangers. This is accomplished when the temperature profile
of the evaporating and condensing mixture matches that of heat-source and heat-sink
fluids in the counter flow desorber and absorber (Mulroy 1993). This is shown in Figure
3-6, which illustrates the Carnot and Lorenz cycles operating with the same heat transfer
fluid temperature profiles at two different source temperatures.

24

Figure 3-6: Irreversibilities in heat transfer processes (modified from Mulroy 1993).
The Carnot cycle refers to a pure refrigerant, for which the evaporation and
condensing process takes place at a constant temperature for a fixed pressure. In
comparison to the Carnot cycle, the Lorenz cycle refers to the evaporation and
condensation process of a mixture, with the temperature glide clearly evident during
these phase change processes. If the profiles of the heat transfer fluid (HTF) flowing
through a heat exchanger in counterflow is modeled with a non-zero slope, the shaded
areas approximate the heat transfer irreversibilities. It can be seen that the irreversibilities
are significantly lower for the Lorenz cycle due to the temperature glide. Furthermore, if
the temperature of the heat source is increased for the same operating cycle, it can be
seen from Figure 3-6 that the temperature profile of the HTF changes. The new profile
can be matched by simply adjusting the mass concentration when using a binary mixture

25
(e.g. to a value of 1 shown in Figure 3-5). A change in the overall concentration of the
mixture circulating through the cycle results in a change of the vapor pressures and

densities at a given temperature, and therefore in a change of the capacity of the system.
It also shows that for a capacity adjustment, the heat transfer irreversibilities remain
nearly constant for the Lorenz Cycle, while they increase notably for the Carnot Cycle. It
is important to note, however, that the saturation temperatures generally do not vary
linearly as a function of enthalpy for zeotropic mixtures, so that Figure 3-6 can be
regarded as a simplification.
Nevertheless, the temperature glide and capacity control are two key features of a
mixture-based cycle. The temperature glide can be used to reduce heat transfer
irreversibilities when the source or sink fluids are not approximated as reservoirs, i.e.,
they have temperature profiles with non-zero slopes through the heat exchange processes.
Furthermore, since the temperature glide can be adjusted over a wide range of heatsource and heat-sink temperatures simply by varying the concentration of` the mixture,
the capacity of the system can be adjusted for peak and non-peak operation.

3.3

Cycle Model of an Organic Rankine Cycle with Solution Circuit

In order to quantify the performance of the ORCSC, it is necessary to build a


thermodynamic cycle model. To simplify the analysis of the cycle, the following general
assumptions are made:

The pressure drop due to frictional losses through the piping and fittings is
negligible. Therefore, the only pressure drops in the system are across the turbine,
pump and expansion valve.

26

The rich solution exiting the absorber and the weak solution exiting the desorber
are both saturated liquids.

The vapor stream exiting the desorber is a saturated vapor.

There is thermodynamic equilibrium between the vapor and liquid phases during
the absorption and desorption processes.

All the piping in the system is perfectly insulated.

There is no oil present in the cycle.


A simulation model was developed to compute the thermodynamic properties at

each state point. Since the working fluid is a binary mixture, an equation of state (EOS) is
required to obtain extensive properties (enthalpy, entropy, etc.) of the mixture at each
state point. Two equations of state were considered: the correspondence method given by
the Lee-Kesler-Plcker (LKP) (Plcker et al. 1977), and the Wide Range Equation of
State by Kunz and Wagner (Kunz et al. 2010). The LKP EOS is given in a form that is
easily modifiable for several working pairs given appropriate interaction parameters. It
was found to perform well for a Carbon Dioxide / Acetone mixture in studies conducted
by Groll and Radermacher (1994). However, the LKP EOS was found to have several
limitations as listed below:

Inability to calculate fluid properties accurately when given a two-component


mixture in the two-phase region. This is particularly important when fixing the
expander and expansion valve outlet states in the ORCSC cycle.

Inability to calculate fluid properties accurately in the superheated or subcooled


regions. This necessitated the assumption of either saturated liquid or saturated
vapor at each state point.

27

Limitations in the range of pressures for which property data were available
(accuracy was limited beyond 70 bars for the CO2 / Acetone mixture).

Limitations in the range of concentrations for which property data were available
(accuracy was limited outside the 0.15-0.6 [kgCO2/kgmixture] range for the CO2 /
Acetone mixture).

Limitations in the range of temperatures for which property data were available
(accuracy was limited beyond 150 C for the CO2 / Acetone mixture).

Due to these limitations, it was found that the cycle performance was significantly overpredicted by the LKP EOS.
The Kunz and Wagner EOS was found to have a greater flexibility of application,
particularly with regard to the range of temperatures, pressures and concentrations for
which useful vapor-liquid-equilibrium (VLE) data were available. Furthermore, the Kunz
and Wagner EOS is integrated with Refprop 9.0 (Lemmon et al., 2012), allowing for easy
retrieval of the thermodynamic properties. The details of the thermodynamic model are
given below.

3.3.1

Mass Balance

The mass balance for the ORCSC relates the mass flow rates of the rich solution,
mr , weak solution, mw , and the vapor stream mv .
m=
mw + mv
r

( 3-4 )

Based on the mass concentrations of the rich and weak solutions, the following mass
balance may be obtained:

28

=
w mw + v mv
r mr

( 3-5 )

A circulation ratio is defined as the ratio of the mass flow rates of the rich solution and
vapor:
CR
=

3.3.2

mr v wk
=
mv r wk

( 3-6 )

Energy Balance

With reference to Figure 3-3, the heat capacity of the desorber can be calculated using the
specific enthalpies evaluated at state points 1, 5 and 6:
=
QD mv ( CR 1) h6 + h1 CRh5

( 3-7 )

Similarly, the heat capacity of the absorber can be calculated using the specific enthalpies
evaluated at state points 2, 3 and 8:
QA= mv h2 + ( CR 1) h8 CRh3

( 3-8 )

The expansion across the expansion valve from state point 7 to 8 is assumed to be an
isenthalpic process:
h7 = h8

( 3-9 )

The energy balance across the internal heat exchanger gives the following equation:
mv CRh5= mv CRh4 mv ( CR 1)( h7 h6 )

( 3-10 )

Based on the isentropic turbine efficiency, turb , the enthalpy at the turbine outlet can be
calculated as follows:
h2 = h1 turb (h1 h2 s )

The power output from the turbine is given by:

( 3-11 )

29
=
Wturb mv ( h1 h2 )

( 3-12 )

Similarly, based on the isentropic pump efficiency, pump , the enthalpy at the pump outlet
can be calculated as follows:
=
h4

h4 s h3

pump

+ h3

( 3-13 )

The power input to the pump is given by:


W
=
mr ( h4 h3 )
pump

( 3-14 )

Finally, the equations used for the thermal (first law) and second law efficiencies are
given below:

thermal =

Wturb W pump

sec ond law =

WD

( 3-15 )

thermal

( 3-16 )

Tsource
Tsink

Note that Tsource and Tsink refer to the temperatures of the heat-source and heat sink
temperature, respectively, assuming that they can be approximated as constant
temperature reservoirs. The heat-source and heat-sink temperatures cannot be reached
because of an incomplete heat transfer in the heat exchangers. Given that the absorber
and desorber incorporate both heat and mass transfer, and the fact that the working fluid
may traverse through the subcooled, two-phase and superheated regions in these heat
exchangers, a traditional definition of heat exchanger effectiveness may not be applied.
Consequently, pinch-point temperatures were used to predict the heat exchanger outlet
temperatures. For the internal heat exchanger, however, the weak solution and rich

30
solution streams remain in the liquid phase, and the traditional heat exchanger
effectiveness definition is used.
With reference to Figure 3-3, the following steps were followed to obtain the
requisite thermodynamic properties at each state point:

State point 6 is set by assuming a saturated liquid mixture at the desorber outlet
(quality of zero), desorber outlet temperature, and a weak solution concentration at
the desorber outlet. Since state points 6 and 1 are in equilibrium in the two-phase
region, providing a temperature and the weak solution concentration leaving the
desorber fixes the state. The vapor concentration is returned as part of the outputs.
Note that a pinch-point is set between the heat source inlet temperature and
maximum working fluid temperature:
Tsource = T1,6 + Tpinch
T1,6 =Tsource Tpinch

( 3-17 )

Initially, this pinch-point temperature is set to 10 K.

State point 1 is set by assuming equilibrium with state point 6. Given the twocomponent two-phase condition that exists at the desorber, only two intensive
properties need to be given to fix state point 1. These may be the same temperature
and pressure as state point 6. The vapor concentration is returned by assuming it to
be a saturated vapor in equilibrium with the saturated liquid weak solution.

State point 3 is set by initially assuming a saturated liquid mixture at the absorber
outlet, absorber outlet temperature, and a rich solution concentration. This gives the
saturation pressure as one of the outputs. Using the saturation pressure, the saturated
condition may be relaxed, and the state point can be adjusted to incorporate some
subcooling at the same saturation pressure. Once again, a pinch-point temperature is

31
applied between the heat sink inlet temperature and minimum working fluid
temperature.
T3, sat = Tsin k + Tpinch
+Tsub T3 = Tsin k + Tpinch

( 3-18 )

Initially, the pinch-point and subcooling temperatures are set to 5 K each.

State point 4 is set by assuming the same high side pressure returned by state point 6,
the rich solution concentration as stated above, and the definition of isentropic
efficiency given by equation ( 3-13 ).

State point 7 is set by assuming a saturated liquid mixture at the internal heat
exchanger outlet, weak solution concentration as given by state point 6, and the same
high side pressure given by state point 6.

State point 5 may be calculated by using the high side pressure given by state point 6,
rich solution concentration calculated by the mass balance, and the enthalpy given by
the internal heat exchanger energy balance in equation (7). Note that an internal heat
exchanger effectiveness is defined based on fluid enthalpies:
=

mh ( hh ,i hh ,o )
Q
=
Qmax m ( hh ,i hc ,i )

min

( 3-19 )

Initially, the heat exchanger effectiveness is set to 0.95.

State point 8 may be fixed by assuming the same enthalpy as state point 7, weak
solution concentration as given by state point 6, and the same pressure low side
pressure given by state point 3.

32

State point 2 is set by assuming the low side pressure set by state point 3, the
concentration and the entropy given by state point 1, and the definition of isentropic
efficiency given by equation ( 3-11 ).

By combining the above steps to calculate the thermodynamic properties with the cycle
mass and energy balance equations, the properties at each state point as well as an overall
solution to the cycle model is achieved.

3.4

Organic Rankine Cycle with Solution Circuit Model Results

This section presents some of the key results from the thermodynamic cycle
model. Table 3-1 summarizes some of the key assumptions mentioned above. Unless
otherwise stated, these assumptions were applied when generating the results described in
this section.
Table 3-1: Key assumptions used in the thermodynamic model.
Description
Condenser, absorber outlet subcooling
Temperature difference between heat sink and condenser/absorber
outlet
Heat sink temperature
Temperature difference between heat source and evaporator/desorber
outlet
Regenerator/internal heat exchanger effectiveness
Pump isentropic efficiency
Expander isentropic efficiency
Negligible pressure drop in lines, separators, and heat exchangers
Negligible heat loss in lines, mixer, separator, pumps, and expander
Complete separation of liquid and gas phases

Value
1. 5 C
2. 5 C
3. 20 C
4. 10 C
5. 0.95
6. 0.6
7. 0.8
8.
9.
10.

Figure 3-7 shows that for a fixed source temperature, the ORCSC has an ideal
concentration pairing which maximizes the efficiency. This means that for a chosen rich

33
solution concentration, there exists an optimal weak solution concentration for which the
cycle performs the best. This is further illustrated in Figure 3-8. Furthermore, as expected,
the trend shows that higher rich solution concentrations lead to higher efficiencies.
However, there are other tradeoffs such as high working pressures that may constrain the
choice of concentration pairings.

Figure 3-7: Impact of ORCSC concentration pairings on efficiency for a source


temperature of 100C and a sink temperature of 20C.
Figure 3-8 clearly shows that there exists an optimal Circulation Ratio for which
the cycle performs the best. Based on equation ( 3-6 ), it is clear the Circulation Ratio is
analogous to the concentration pairing, and therefore confirms the results shown in
Figure 3-7. The Circulation Ratio, however, is an important physical parameter that is
essential to the control of an ORCSC system. It fixes the relative mass flow rates of the
rich solution and vapor, and therefore fixes the relative speeds of the pump and expander
in the system. By extension, it fixes the concentration pairing of the system. Therefore

34
the term Circulation Ratio is used in this text as opposed to concentration pairing,
particularly when discussing the physical ORCSC system and the experimental study.

Figure 3-8: Variation of overall Second Law Efficiency as a function of Circulation Ratio.
Figure 3-9 illustrates the effect of rich solution concentration on overall cycle
Second Law Efficiency for a Carbon Dioxide / Acetone working pair. Note that the
chosen points have been optimized with respect to weak solution concentration using
Engineering Equation Solver (EES) (Klein, 2012). From the results shown in Figure 3-7,
and from a review of absorption studies in the literature, an expected trend would show
that higher rich solution concentrations lead to higher the efficiencies. The trend in Figure
3-9 matches this expected outcome for high concentration values. Note that a peak value
may not occur in the high concentration region because the two-phase concentration
range shrinks as the mixture critical point is approached. This means that at extremely
high concentrations; it may not be possible to operate a solution circuit. Furthermore, an
interesting overall result exists in which the peak efficiency is achieved at extremely low

35
concentration values. This means that a large fraction of the absorbent, in this case
Acetone, participates in the expansion process. One reason for this phenomenon, shown
in Figure 3-10, is that at extremely low concentrations the absorbent, Acetone, is a better
working fluid than the absorbate, Carbon Dioxide as pure fluid. Therefore, at extremely
low concentrations, the highest efficiencies are achieved because the component with
better thermodynamic properties dominates the mixture, while the overall cycle
simultaneously takes advantage of the internal regeneration intrinsic to the solution
circuit. Achieving this optimal efficiency in the low-concentration region, however, is a
practical challenge. Extremely tight tolerances in solution concentrations are required to
realize these benefits. Nevertheless, the peak efficiencies that occur at extremely low
concentrations have not been investigated in the literature, and merits further
investigation in the future.

Figure 3-9: Impact of rich solution concentration on efficiency for various source
temperatures. Plot is optimized for weak solution concentration.

36

Figure 3-10: Second Law efficiency as a function of source temperature for a basic ORC.

Figure 3-11: Second Law efficiency as a function of source temperature for an ORC with
internal regeneration at the expander outlet.

37

Figure 3-12: Second Law efficiency as a function of source temperature for the ORCSC.

Figure 3-10 and Figure 3-11 compare the optimized performance of various
working fluids for a basic ORC and an ORC with internal regeneration, respectively,
using the parameters listed in Table 3-1. Figure 3-11 shows that the addition of a
regenerator results in better performance for dry working fluids, such as R245fa and
pentane, at high temperatures because the availability in the expander exhaust stream is
not wasted. However, wet fluids like water and ammonia show no improvement because
the expander exhaust temperature is not significantly higher than the pump discharge
temperature.
Figure 3-12 shows the optimum performance of an ORCSC with Ammonia /
Water and CO2 / Acetone working fluid pairs at two different concentration ranges: at
extremely low absorbate concentrations (less than 1% by mass), in which case a large
fraction of saturated water vapor or Acetone vapor participates in the expansion process;
and at high concentrations (above 50% by mass), where only small amounts of saturated

38
Water vapor or Acetone vapor participates in the expansion process. However, at a
200 C source temperature for the CO2 / Acetone working pair, it was not possible to
operate at rich solution concentrations above 10% because the two-phase concentration
range shrinks as the mixture critical point is approached. In this case, the highest possible
concentration of 10% was used.
Note that in addition to the assumptions shown in Table 3-1, there were no
restrictions on vapor quality at the expander exhaust, expander pressure ratios,
Circulation Ratios, amount of charge required in the system for a given capacity, or
specific volume of the working fluids at the expander discharge. Therefore, the results
presented in Figure 3-10, Figure 3-11 and Figure 3-12 are highly idealized and only
represent what is possible in the thermodynamic design space. For example, Water
requires a pressure ratio of 296 for optimum efficiency at 200C. It also has a specific
volume at the expander discharge of nearly 50 m3/kg at the same design point. Designing
an expansion machine with multiple stages to accommodate the high pressure ratios,
along with the cost of large components to accommodate the low density renders Water
an impractical working fluid. Several of these practical concerns are mitigated in the
ORCSC system.
The ORCSCs with high concentrations of ammonia or CO2 are generally less
efficient than traditional ORCs using pure ammonia or CO2 respectively. In these cases
the temperature glide is detrimental to the Second Law efficiency because it results in
higher irreversibilities against the assumed constant temperature heat source. A real heat
source fluid will begin to cool as it heats the cycle working fluid. This could severely
limit the maximum evaporating temperature of a traditional ORC as the heat source

39
stream reaches a pinch point against the liquid to two-phase transition of the working
fluid. In contrast, the ORCSC may be able to reach higher average working fluid
temperatures by being better able to match the temperature profile of the heat source
stream with its two-phase temperature glide. A similar argument can be made between
the working fluid and the heat sink fluid. Exceptions occur for ammonia-water above
source temperatures of 160 C and for CO2-acetone at 200 C, where the CO2
concentration is relatively low. The ORCSCs with low concentrations of ammonia and
CO2 do not perform as well as traditional ORCs with pure water and pure acetone
respectively (Woodland et al. 2012).
Figure 3-13 shows the desorber capacity along with the pump power input as a
function of the Circulation Ratio for a fixed rich solution concentration. Note that the
desorber capacity and pump power input have been normalized for a 1 kW turbine design.
As the Circulation Ratio increases, the amount of CO2 in the weak solution is increased
and the pump input power increases. More important, however, is the fact that the
desorber capacity decreases as the Circulation Ratio is increased. This illustrates a
solution for capacity control in this system during non-peak loads, the Circulation Ratio
may be increased either by a concentration adjustment or, to lesser extent, by varying the
relative speeds of the pump and expander to accommodate a smaller heat input at the
source (Krishna et al. 2011).

40

Figure 3-13: Variation of component capacities as a function of Circulation Ratio.


Figure 3-14 indicates that the system pressures increase considerably as the
Circulation Ratio is increased. Therefore, with reference to Figure 3-8, it is clear that
there is a tradeoff between the efficiency and the operating pressures seen in the system.
With reference to Figure 3-13, it can also be deduced that there is a relationship between
capacity control and the system pressures. As the capacity is lowered by increasing the
Circulation Ratio, the system sees higher pressures. Figure 3-15 shows the linear
variation of the Pressure Ratio with the Circulation Ratio. This is an important design
consideration, particularly with regard to the expansion machine. Depending on the
desired capacity range of the system, it would be advisable to design and optimize the
expander for the corresponding Pressure Ratio.

41

Figure 3-14: Variation of High-side system Pressure as a function of Circulation Ratio.

Figure 3-15: Variation of Pressure Ratio as a function of Circulation Ratio.


Based on the results shown in Figure 3-10 and Figure 3-12, it may be tempting to
conclude that the ORCSC does not offer sufficient benefits, and a basic ORC with the
best performing working fluid would be the ideal choice due to a combination of

42
simplicity and performance. However, it is important to note that several practical
considerations should be weighed together with the thermodynamic results presented
above. For example, the optimum efficiency for a basic ORC with water as the working
fluid and a source temperature of 200 C is significantly higher than the theoretical
efficiencies of the ORCSC. However, the use of water requires a low vapor quality and
low density at the expander exhaust, vacuum pressure condensation, and extremely high
expander pressure ratios. High pressure ratios require multiple expansion stages. The subatmospheric pressure makes air leakage into the system a challenge. The low working
fluid density is a capacity concern, requiring large diameter piping and a large expander
to achieve a capacity comparable to denser working fluids. Therefore, despite the
promising efficiency of water as a working fluid for the source temperatures considered,
the practical issues associated with its use may be prohibitive.
The hydrocarbons acetone and pentane are very efficient, but they are highly
flammable and pose challenges for sealing materials. This leads to the widely accepted
view that R245fa is perhaps the best working fluid in an ORC. However, the ORCSC
offers some potential practical advantages over an ORC with regeneration using R245fa
as the working fluid that may outweigh the added complexity of the system. The ORCSC
using ammonia-water as the working fluid at high concentrations may have marginally
lower efficiencies than a R245fa ORC with regeneration, but the use of a zeotropic
mixture in the ORCSC provides the ability to use the temperature glide to match the
temperature profiles of the source and sink fluids. As shown in Figure 3-13, the zeotropic
mixture also allows for capacity control simply by changing the concentration of the
working fluid mixture. This may lead to higher overall system efficiencies when coupled

43
with sources that have varying heat input temperatures or loads, such as waste heat
streams from diesel engines, power plants or other industrial processes. The use of a
Carbon Dioxide / Acetone working pair is difficult to justify from a performance
standpoint; its main advantage lies in the fact that aside from CO2 being an inexpensive,
natural, non-flammable refrigerant, its high volumetric heat capacity allows for the use of
smaller components which may lead to significant weight savings in mobile applications.
The ORCSC is not without its challenges. The highest efficiencies for the ORCSC
using Ammonia-Water require significantly low vapor quality in the expander exhaust.
This is because the desorber outlet state point is modeled as a saturated state, with liquid
and vapor existing at the same temperature and pressure (with differing concentrations).
Therefore, since the vapor is at a saturated condition at the expander inlet, the expansion
process has a tendency to generate significant quality at the outlet condition. This is not
an issue with the Carbon Dioxide / Acetone working pair because both Carbon Dioxide
and Acetone are dry working fluids. Therefore, the shape of the vapor dome dictates that
the expander outlet condition will not fall into the two-phase region. Control related
issues, which become significantly more complex with the ORCSC compared to the
traditional ORC, have not been investigated in detail in the literature, and require
rigorous examination.
Thus, a truly optimal choice of ORC and working fluid may be highly
application-specific. Only two working fluid pairs were studied out of a broad range of
possible binary mixtures. More binary mixture data is needed to study the wide range of
working fluid combinations that may be possible with the ORCSC. Appropriate system
selection, therefore, requires a much more thorough investigation of the tradeoffs

44
between efficiency, ability to match the heat source temperature profile, cost of
implementation, and size constraints. Investigating these tradeoffs, however, requires
experimental study. In this regard, an ORCSC experimental test stand was constructed in
order to provide insight into the feasibility of the technology, and further develop the
results of this study.

45

CHAPTER 4. DESIGN OF EXPERIMENTAL TEST SYSTEM

An experimental test setup (breadboard system) was constructed to serve as a


proof of concept for ORCSC technology. The system was fabricated largely using offthe-shelf parts harvested from the Liquid-Flooded Ericsson Cooler experimental test
setup (Hugenroth 2006) at the Herrick Laboratories. Since the breadboard system was
constructed before the computer modeling was complete, several estimates pertaining to
system design were necessary. The timing for the design and construction of the system
were largely dictated by needs of the project sponsor. In addition to proof of concept
testing, the breadboard system was intended to validate the results of the thermodynamic
cycle model, and to gain design-related experience.
4.1

System Sizing

The experimental breadboard system was designed for a nominal 1 kW power


output. Table 4-1 shows the key input parameters used in the simulation model. The
values reflect initial expected performance of the equipment, and were used as a basis for
system design and component selection.

46
Table 4-1: Sample set of input parameters used as a basis for system design.
Symbol Value
Description
Unit
CR
5
Circulation Ratio
[-]
,
0.7
Isentropic turbine efficiency
[-]

0.5
Mechanical pump efficiency
[-]
80
Temperature level of heat-sink
[ C]

20
Temperature
level
of
heat-source
[ C]

5
Pinch Point temperature
[ C]
0.21
Weak Solution Mass Concentration
[-]

0.04
CO2 vapor mass flow rate
[kg/s]

1.288
Heat Capacity ratio of the working fluid
[-]
Based on the data in Table 4-1, the conditions at each state point in the cycle were
calculated as shown in Table 4-2. Note that these initial calculations were performed
using the LKP EOS and contain significant sources of error as described in Section 3.3.
However, these were used as the initial design basis and are presented here to illustrate
the parameters taken into account when constructing the experimental bread board
system.
Table 4-2: Sample property values at each state point in the ORCSC cycle using the LKP
EOS (refer to Figure 3-3).
State
T [ C]
p [bar]
h [kJ/kg]
s [kJ/kg/K]
[kg/m3]
point
1
75.00
29.85
579.0
2.4
49.7
2s
49.41
21.68
560.2
2.4
39.0
2
41.57
21.68
552.2
2.3
40.4
3
25.00
21.68
218.7
1.2
829.9
4
25.68
29.85
219.7
1.2
829.9
5
59.49
29.85
270.4
1.3
769.2
6
75.00
29.85
283.2
1.3
707.2
7
25.68
29.85
219.9
1.6
778.2
8
25.64
21.68
219.9
1.6
778.2
Based on the sample data given in Table 4-1 and Table 4-2, the predicted
capacities, flow rates and cycle efficiencies using the LKP EOS are shown in Table 4-3.

47
These were used as the primary design parameters when developing the ORCSC
experimental breadboard system.
Table 4-3: Capacities, flow rates and cycle efficiencies used as a design basis for the
ORCSC.
Description

Symbol

Value

Capacities
Desorber Capacity
14.4

Absorber Capacity
13.53

Internal Heat Exchanger Capacity


10.13

Work input pump


0.39

Work output turbine


1.07

Mass Flow Rates and Mass Concentration


Vapor flow rate
0.04

Weak solution
0.16

Rich solution
0.20

Rich Solution Concentration


0.37

Cycle Efficiencies
Thermal Efficiency
4.7

Carnot Efficiency
17.0

2nd Law Efficiency


27.9
2

Unit
[kW]
[kW]
[kW]
[kW]
[kW]
[kg/s]
[kg/s]
[kg/s]
[-]
[%]
[%]
[%]

It is important to note that the output values listed in Table 4-2 and Table 4-3
reflect the input parameters presented in Table 4-1. Therefore, these outputs may be
considered as general design parameters. For a different set of input data, the output data
can vary significantly. Furthermore, due to the inaccuracies in the LKP EOS which were
discovered after the construction of the test setup, several of the design values are
themselves inaccurate. Nevertheless, these were the parameters taken into account when
constructing the experimental bread board system.

48
4.2

System Design and Layout

This section describes the design and layout of the various components in the
system. Figure 4-1and Figure 4-2 show a more detailed evolution of the system schematic
presented in Figure 3-3. The figures include the flow paths of the source, sink and
working fluids through the system. Additionally, all valves utilized to operate the test
stand, charge and drain the system, measure the concentration (, ), as well as those that
provide for maintenance of the system are shown. The locations of measurement

equipment in the test loop, including pressure transducers and thermocouples, are also
indicated. The placement of safety instruments such as pressure relief valves (PR) and
pressure switches (PS) are also shown. As seen in Figure 4-2, two major components
were added to the system when compared to the cycle schematic presented in Figure 3-3,
namely a separator and a receiver. The separator is a requisite addition in order to provide
additional volume for the desorption process that may not be available in the desorber
itself. The receiver, on the other hand, is primarily used to maintain a liquid head and
prevent cavitation inside the pump. Note that based on preliminary experiments, it was
decided to include a secondary loop, shown in Figure 4-1, to provide better temperature
control of the heat input. This would enable a better assessment of system performance,
particularly with respect to cycle efficiencies and desorber performance. Therefore, a
secondary loop that exchanged utility steam available in the testing facility with an
aqueous ethylene glycol solution was constructed for this purpose.

49

Figure 4-1: Detailed schematic diagram of the secondary loop for heat input temperature
control.

Figure 4-2: Detailed schematic diagram of the primary ORCSC system.

50
Based on the schematic diagram shown in Figure 4-2, a CAD model of the test
stand was drawn in ProEngineer as shown in Figure 4-3 before the setup was actually
built. The goal of this preliminary design was to decide on a suitable arrangement of the
major components on the relatively small experimental test table. It was also used to
minimize the required piping between the components for the rich solution and vapor line
and thereby minimize pressure losses. In contrast, larger piping distances for the weak
solution circuit were tolerated due to the lower priority of pressure losses in this circuit,
and to allow for the best possible piping for the rich solution and vapor lines. The lower
emphasis on pressure losses in the weak solution circuit is due to the fact that an
expansion occurs across the expansion valve in any case during normal operating
conditions. If a hydraulic motor was used in place of an expansion valve to recover the
potential of the fluid stream, the pressure losses in the weak solution circuit would need
to be minimized.
Figure 4-3 illustrates the arrangement of the major components of the system and
the piping connections that link them. Aside from refining the equipment layout, the
drawing was used to estimate space availability between the components to insert
equipment such as valves, pressure transducers, and thermocouples required for system
operation. It can be seen that the piping for the rich solution (blue) and vapor line (yellow)
consist of the shortest distances, whereas the weak solution lines (green) are of longer
distances. The reasoning behind the different elevations of certain components, especially
the heat exchangers, is discussed in the following sections. Figure 4-5 and Figure 4-6
show the completed ORCSC load stand and the secondary heat input load stand,
respectively.

51

Figure 4-3: Preliminary CAD representation of ORCSC system arrangement.

52

Figure 4-4: Top view of the primary ORCSC experimental load stand.

Figure 4-5: Front view of the primary ORCSC experimental load stand.

53

Figure 4-6: Secondary heat input loop load stand.


4.3

Design and Selection of Major Components


4.3.1

Pump

The selection of the pump is based on the outputs from the simulation model for
several sets of input parameters, as well the specific properties of the working fluid. The
nominal criteria are listed below:
a. = 20 ;
b. = 0.20

= 35

= 3.9 14.5 /

54
c. Sealing and construction materials compatible with Acetone as well as Carbon
Dioxide.
d. Capable of pumping a low viscosity fluid, such as the Carbon Dioxide / Acetone
mixture.
Note that these values correspond to a single set of input data and do not cover all
possible operating conditions required for a thorough experimental study. For example,
when doubling the circulation ratio, the mass flow of the rich solution roughly doubles as
well. Increasing the absorbate concentrations has a profound effect on operating pressures.
Pressure losses, which were assumed to be negligible in order to set acceptable boundary
conditions for the simulation, can increase the required pressure lift across the pump as
well. Therefore, the chosen pump must be able to handle elevated discharge pressures
and mass flows. However, to achieve a volume flow rate of about 3.9 gpm and a pressure

lift of 15 bars is challenging for a small-scale experimental application without resorting


to multiple stages and/or machines. In order to eliminate any system leakage, particularly
due to the flammability of the mixture, the pump selection was narrowed down to a few
possibilities. Table 4-4 qualitatively compares potential pumps to be used for the
experiments:
Table 4-4: Qualitative comparison for different pump types for an ORCSC application.
Type of pump
Sealing
Mass flow Head Viscosity Cavitation Price
Gear / screw
++
+
++
-
+
Centrifugal
++
++
-+
-Diaphragm
++

-Plunger

++
++

+
+
Legend: ++ very good; + suitable; would work; - poor; -- knock out criteria
As can be seen from Table 4-4, a gear/screw pump would potentially satisfy all
criteria. This type of pump can be driven by a magnetic coupling, so that no rotating shaft

55
is in contact with the working fluid. Also, these pumps can provide the required pressure
lift with satisfactory mass flows using a single stage machine. Gear/screw pumps require
a lubrication film to avoid a direct metal-to-metal contact in between the two rotating
gears. These pumps are usually installed in hydraulic applications with high viscosity
fluids; however the low viscosity Carbon Dioxide / Acetone mixture does not provide the
necessary lubrication for this purpose. Therefore, it was determined that gear/screw
pumps would not be appropriate for this application.
Magnetically driven centrifugal pumps are limited by the maximum differential
pressure produced per stage. For the required pressure lift in an ORCSC application, the
manufacturers stated that at least four stages in series were required. Thus, a multi-stage
customized pump would exceed the available space in the test rig and would be
prohibitively expensive. Furthermore, centrifugal pumps are extremely prone to
cavitation due to the high relative velocities between the surface and working fluid.
Diaphragm pumps may also be considered for this application provided that the
diaphragm is compatible with Acetone. A key feature of diaphragm pumps is that only
the diaphragm is in contact with the fluid. As a result, one of the advantages of
diaphragm pumps is that they provide excellent leak protection. During the course of the
pump selection, however, it was found that the only diaphragm material compatible with
the mixture was that made of Polytetrafluoroethylene (PTFE). Diaphragm pumps are
normally designed for smaller mass flows, and finding a diaphragm pump capable of
exceeding a flow rate of 4 gpm was extremely challenging. The pumps that met the mass
flow requirements were found to be extremely large and expensive. Based on the

simulation results, it was determined that a reasonably priced diaphragm pump would

56
limit the range of circulations ratios that could be tested and therefore would not be an
ideal selection.
To meet the requirement for high mass flows, elevated inlet/discharge pressures,
as well as large pressure lifts, plunger pumps with multiple pistons running out of phase
in parallel were the sole design matching all requirements. In comparison to a diaphragm
pump, a plunger pump with three plungers in parallel theoretically has lower pressure
fluctuations leading to reduced vibrations. In this regard, a CAT PUMP Model 1051.CO2
designed for pumping pure Carbon Dioxide was found to fulfill all the requirements. The
primary pump specifications are listed in Table 4-5. The specifications guaranteed that
the pump would handle a wide range of operating conditions, and would be effective for
the purposes of a parametric study.
Table 4-5: Primary specifications of CAT PUMP MODEL 1051.CO2.
35 bars (500 psi)
maximum
maximum
125 bars (1800 psi)

38 l/m (10 gpm)


maximum

Figure 4-7: Sectional view of CAT PUMP MODEL 1051.CO2.

57
Figure 4-7 illustrates a sectional view of the pump and is useful to understand the
working mechanism: The crankshaft (3) is located in the crankcase (1) which is filled
with an oil bath and is driven by a motor/belt drive available at the Herrick Laboratories.
The belt drive powers the three plungers (7) in parallel. The pump is connected via the
two manifolds (8) to the suction (larger diameter) and discharge sides (smaller diameter),
respectively. The suction and discharge processes are regulated by the check valves (10).
In comparison to a magnetically driven pump with no contact between the
working fluid and sealing, the fluid in a plunger pump is in contact with static and
dynamic seals. These seals are O-Rings located on the plunger and check valves,
respectively. Therefore, O-Rings made of Ethylene Propylene Diene Monomer (EPDM),
which were compatible with the Carbon Dioxide / Acetone mixture were chosen. The
pump manufacturer guaranteed an operation time of 2500 hours provided that an
adequate liquid head was maintained on the pump. In case of a loss of head and
consequent pumping of vapor, the seals were expected to fail within minutes. Generally
with dynamic seals, a 100 % leak-proof seal is difficult to achieve. In the event of a
leaking plunger seal, however, the leaked material first accumulates in the crank case.
The crank case itself is sealed from the atmosphere. Therefore, the crankcase may be
regarded as a containment of last-resort in case of a broken plunger seal.
4.3.2

Heat Exchangers

The overarching consideration when selecting a heat exchanger design for the
external heat exchangers (absorber and desorber) was its ability to facilitate mass transfer
as well as heat transfer. Therefore, shell and tube heat exchangers would have been ideal
due to the ample volume provided on the shell side for the absorption and desorption

58
processes. However, finding shell and tube heat exchangers capable of withstanding high
system pressures was challenging. Shell and tube heat exchangers rated for high system
pressures are readily available for large scale industrial applications; however, no vendor
that offered shell and tube heat exchangers at reasonable cost, rated for high pressures
and sized for a small scale laboratory application, was found. Therefore, a compromise
dual coil heat exchanger design from Sentry Equipment (model DTC-CUB/CUC-8-1-1
for the absorber/ desorber and model DTC-CUA/CUB-6-1-1 for the internal heat
exchanger) was selected as shown in Figure 4-8. This design was found to work well for
Groll et al. (1994) in experimental work for the VCCSC cycle. Due to better thermal
conductivity and lower cost, copper tubing for both the inside and outside coils was
chosen over stainless steel. It must be noted, however, that even with a dual coil design,
the heat exchangers have the lowest Maximum Allowable Working Pressure (MAWP)
rating of any equipment in the system (48 bars). All heat exchangers are operated in
counter flow to reach the highest possible outlet temperatures, and thereby maximize
efficiency for the ORCSC cycle configuration.

Figure 4-8: Dual Coil Heat Exchangers selected for use in the ORCSC.

59
The primary purpose of the desorber is to facilitate heat exchange between the
heat source and the system. Utility steam available at the Herrick Laboratories was
initially used to simulate a high temperature reservoir. However, due to pressure
limitations of several pieces of equipment, it was necessary to lower the temperature of
the heat source. Temperature control was not available on the steam line, and a secondary
loop (Figure 4-1) was built for this purpose as discussed earlier. The high temperature
glycol flows through the inner coil of the desorber, which has a inch (1.27cm) outer
diameter. Temperature and flow control of the glycol loop is achieved by a combination
of a pressure regulator and balancing valves on the steam line, bypass valves on the pump
and a three-way electronic bypass valve on the steam/glycol heat exchanger (Figure 4-1).
The rich solution enters the desorber in counter flow configuration from the bottom. Note
that the rich solution comes from the outer coil of the internal heat exchanger, which has
a inch (1.27cm) outer tubing diameter, and enters the outer coil of the desorber which
has a 1 inch (2.54cm) outer diameter. This corresponds to a 25% increase in flow area,
and is analogous to an expansion when entering the desorber. This is advantageous
because it facilitates desorption of the carbon dioxide. Also, by entering the desorber
from the bottom, the rich solution flows upward and takes advantage of the buoyancy
forces to avoid clogging the flow.
The purpose of the absorber is to facilitate heat exchange between the system and
the heat sink. Since absorption is an exothermic process, it is necessary to cool the
mixture solution. A utility water stream was used to simulate a low-temperature reservoir.
Flow control is achieved by the use of balancing valves. Temperature control was not
deemed necessary since a natural low-temperature reservoir would be the ambient

60
environment. Before entering the absorber, the vapor and the weak solution streams are
mixed. In contrast to the desorber, the mixture enters the absorber from the top and leaves
at the bottom. By using this flow configuration, vapor ascends to the top so that only
liquid leaves the absorber. The exit of the absorber is located at the same elevation as the
inlet of the receiver. The tubing between the two devices is at the same level, and thereby
avoids any bubbles accumulating and clogging the flow. The elevation levels of the other
devices are fixed and necessitate the relatively high elevation of the absorber. For
example, the elevation of the pump is fixed due to the available belt drive powered by a
motor located underneath the table. A liquid head needs to be maintained on the pump; so
the outlet of the receiver is at the same level as the suction side of the pump. Therefore,
the receiver is at a certain elevation above the table, which necessitates the absorber
being even higher.
4.3.3

Expander

In general, of the major mechanical components used in the ORCSC, heat


exchangers and pumps have been widely studied, and several commercial products exist
in the market. However, expansion turbines used in ORC or ORCSC technology have not
been widely studied, and few commercial designs are available. Note that the term
expansion machine or expander is used to convey that a positive displacement
machine such as a scroll or screw machine is being used, as opposed to a dynamic
machine such as a turbine found in a steam power plant or an aircraft engine. Positive
displacement machines rely on volumetric expansion of the gas to generate shaft power,
whereas turbines rely on velocity conversion to do the same. Positive displacement

61
machines have the distinct advantage of achieving high efficiencies at small scales for
distributed power generation or waste heat recovery.
In order to be effective for ORCSC experimental study, the expander needs to be
flexible enough to handle different operating conditions such as varying mass/volume
flow rates and inlet and outlet temperature and pressure conditions that lead to different
expansion ratios. Unfortunately, positive displacement machines are usually designed for
a specific expansion ratio that is fixed by the geometry of the expansion machine. This
problem can be mitigated by using a design that includes suction and discharge valves to
control the flow of gas into and out of the expander. However, to operate efficiently, the
volume trapped in the expansion chamber must be nearly zero when the discharge valve
is opened. This is because any gas remaining in the chamber will re-compress during the
intake stroke. This reduces both the adiabatic and volumetric efficiencies of the expander.
Previous work by Hugenroth (2006) and Lemort et al. (2008) has shown that
scroll compressors modified to run in reverse may work successfully in this application.
The scroll expander simply uses what is normally the compressor discharge as the
expander inlet. The pressure differential from the expander inlet to outlet causes the fluid
to move in the opposite direction through the scroll wraps, expanding in the process, and
rotating the shaft in the direction opposite to that of compression. Initially, a Sanden
model TRS-105 shown in Figure 4-9, typically used with R134a refrigerant in automotive
applications and used successfully by Hugenroth (2006), was incorporated in the ORCSC
system. However, preliminary tests showed that the Sanden scrolls designed
displacement volume of 104.8 cm3 was far too large for the test conditions chosen for the

62
ORCSC. Therefore, the scroll expander was forced to run at extremely low speeds that
were not sustainable for efficient or reliable operation.

Figure 4-9: Sanden TRS-105 scroll compressor housing.


Due to the unsuitability of the Sanden scroll compressor, a prototype rotary spool
expander, shown in Figure 4-10, produced by Torad Engineering was used as a
replacement. The Torad rotary spool expander has a unique design (see Figure 4-11) that
combines rotary motion, analogous to a rolling piston machine, with low friction sealing
technology. The working fluid enters the rotary spool expander through the inlet port.
The fluid expands against the rotating gate, providing positive torque and producing
positive work output. The rotary spool expander delivers two expansion cycles per
revolution, allowing the trapped fluid in the intermediate working chamber to expand
toward an equilibrium force balance. This attribute, combined with the placement of the
intake and discharge relief ports, allows the machine to operate smoothly. Due to the
design, the forces acting on the rotating components are small compared to other
technologies such as screw and scroll expanders. It also provides higher power density as
compared to screw and scroll expanders. However, without using suction and discharge
valves to control the flow of gas into and out of the expander, the expansion ratio is fixed

63
by the relative location of the inlet and discharge ports. This limits the range of operating
conditions for which the expander is effective. Ideally, fast acting electronic valves such
as those used in automotive fuel injection, or a mechanical design which couples the
valve opening / closing to the relative location of the rotating gate is needed for efficient
and flexible expander operation. However, this was not available is the provided
prototype, and most experimental tests were performed by using the shutoff valves across
the expander as throttling devices.

Figure 4-10: Front view of Torad rotary spool expander housing.

64

Figure 4-11: Design of the Torad rotary spool expander.


4.3.4

Separator

The function of the separator is to effectively divide the fluid that exits the
desorber and divide this flow into a pure vapor stream and a liquid stream. To have
enough volume and surface available for an effective separation process, two hollow high
pressure vessels were assembled in parallel. The mixture exiting the desorber is divided
into two equal flows before entering each separator from the top (see Figure 4-12). From

65
there the fluid empties into the shell where the liquid and gas are separated by gravity. A
dip tube for liquid pickup was fed through the top of the shell and extended nearly to the
bottom of the separator. The number of holes increases from the top to the bottom to
establish equal flow along the dip tube. After the separation process, the vapor is
extracted through a port mounted flush with the top of the separator. The flows from each
vessel combine before entering the expander. Simultaneously, the weak solution exits
each vessel at the bottom, before combining and flowing to the internal heat exchanger.
The separator also contained service valves at the bottom of the shell. These were used
for system evacuation and charging of both liquid and gas. An upper and lower sight
glass was installed in a parallel line along the height of the separator to allow for the
inspection of the contents. There is no flow through the sight glass line due to equal
pressures of the vapor stream and the weak solution stream.

Figure 4-12: Design and layout of separator and sight glass.

66
4.3.5

Receiver

The function of the receiver is to maintain a liquid head on the pump. It is


essentially a storage device elevated above the suction side of the pump to ensure a liquid
level and reduce the risk of cavitation inside the pump. A sight glass was mounted at the
outlet of the receiver to allow for verification that only liquid leaves the receiver and
enters the pump.

Figure 4-13: Receiver and sight glass selected for use in the ORCSC.
4.3.6

Expansion Valves

The sub-cooled weak solution exiting the internal heat exchanger needs to be
expanded to the same pressure as the vapor stream from the expander outlet before
entering the absorber. Therefore, two needle valves in parallel were used to adjust the

67
pressure across the valves. One of the needle valves served as a rough adjustment,
whereas the other needle valve was chosen for precise control.

4.3.7

Instrumentation and Data Acquisition

Figure 4-14 shows a general overview of the data acquisition system. The
following was the primary instrumentation installed on the load stand:

17 T-type thermocouples

1Setra Model 207, 5 OMEGA PX32 Series and 1 OMEGA PX172 Series
pressure transducers

1 Althen Model 01324 torque sensors

Motor speed output from Baldor Vector drives

Three Micromotion coriolis mass flow meters

Figure 4-14: General overview of the data aquistion system.


All instrumentation data was collected by an Agilent 34980A data acquisition
system via a USB/GBIB interface cable. An Agilent VEE program was written to display
and store the data during testing. The signal type from each piece of equipment is detailed
in Table 4-6.

68
Table 4-6: Data acquisition devices and their output signals.
# of Channels
Description
Signal
Thermocouples [T]
17
OMEGA T type thermocouples
voltage
Pressure transducer [p]
5
OMEGA PX32 Series
voltage
1
Setra Model 207
voltage
1
OMEGA PX172 Series
voltage
Flow Meter [mass flows]
1
ampere
F Series (SN 2187659) - measuring
1
ampere
R Series (SN 2203437) - measuring
R Series (SN 2173136) - measuring
1
ampere
Torque Cell [M]
1
01324 Althen Torque Sensor
voltage
Revolutions per minute [RPM]
Baldor 18H Controllers (pump and turbine
2
voltage
drive)
Motor Drives [M]
Baldor 18H Controllers (pump and turbine
2
voltage
drive)
System capacity and heat rejection rates were calculated using the flow and
temperature data across the heat exchangers. Since only one torque sensor was available,
it was installed on the expander. Electric motor speeds were output by the motor
controller. From torque and speed, the shaft power of the expander was calculated.
For additional information regarding data acquisition, including instrument
calibration, please see Appendix A. For information regarding load stand operation,
please see Appendix D.

4.4

Experimental Error and Uncertainty

In order to quantify the accuracy of experimental results, a propagation of error


analysis was performed on the experimental data analogous to that performed by
Hugenroth (2006). The error contribution of each measured parameter to the overall

69
uncertainty of the calculated parameter was determined. The overall uncertainty is
defined as (Kline and McClintock, 1953):
1

j A 2 2
wA =
wz
i =1 zi

( 4-1 )

where wA is the total uncertainty of the calculated quantity, A is the calculated quantity,
zi is the ith measured quantity, and wz is the uncertainty in the measurement of zi .
The heat transfer rate of the heat exchangers was determined from the enthalpy
change across the inlet and outlet conditions:
=
Q m ( hout hin )

( 4-2 )

The total uncertainty for the heat transfer process is:


2

Q
Q=
wm +
wTi +
wPi

m i in=
, out Ti
=
i in ,out Pi

( 4-3 )

Power for the rotating machinery can be calculated from shaft torque and
rotational speed, or by flow rates along with inlet and outlet temperatures and pressures.
The equation for shaft power is given by:

Ws = Tor

( 4-4 )

The total uncertainty is given by:


2

=
wWs

Ws
W

wTor + s w

Tor

( 4-5 )

The thermodynamic power is given by:


=
Wt mv ( hin hout )

( 4-6 )

where:
h = h ( T , P, )

( 4-7 )

70
Note that the concentration is measured by taking a physical sample. Therefore,
the uncertainty in the concentration value is directly attributable to the uncertainty in the
mass measurement for this purpose.
The total uncertainty is given by:
2

wWt = t wmv + t wTi + t wPi + t wi


=
, out Ti
, out Pi
mv
i in=
i in=
i in ,out i

( 4-8 )

The cycle efficiency is given by:

Wnet
Qin

( 4-9 )

The total uncertainty becomes:


2

w =
wml +
wTi +
wTor
w
=
out Ti
p Tori
ml
i in ,=
i e,=
i e, p i

( 4-10 )

where the subscripts e and p correspond to the expander and pump respectively.
EES has a built-in functionality that computes the partial derivatives for the
uncertainty analysis using user specified measurement uncertainties. The measurement
uncertainties are shown in Table 4-7.
Table 4-7: Measurement uncertainties.
T (C)
P
Tor

mv

(kPa)
0.5

(N-m)

(Hz)

4.48 1.0% 1.0% 0.0136 0.0167

71

CHAPTER 5. EXPERIMENTAL RESULTS AND ANALYSIS

An experimental test matrix was developed for the ORCSC in order to prove the
viability of the technology and validate the results from the simulation model. As far as
the equipment allowed, an attempt was made to quantify the performance of all the major
components in the cycle. The results of the test program are presented in this chapter. An
overview of the experiments that were run as well as limitations that were encountered
during testing is discussed.

5.1

Experimental Program Overview

The commissioning phase included shakedown tests that verified equipment


operation, charge inventory and instrumentation response. No data was collected during
this phase. The initial experimental program focused on examining the effect of source
temperatures on system performance while keeping the other variables fixed. Six
independent system parameters are required to fix the operating condition of the system:
the source and sink temperatures, the expander inlet pressure, the Circulation Ratio which
is essentially a direct function of the expander-to-pump speed ratio, the pump speed and
the concentration of the rich solution. Note that the operating methods used to fix these
specific parameters varied to some extent. The tests consisted of 20 experimental runs.
The operating conditions for the experimental tests are shown in Table 5-1.

72

Run
index

P1

Table 5-1: Experimental test matrix.


Pump
CR
Tsource (C)
r

(bar)

speed (RPM)

Tsin k (C)

30

150

0.4

55

18

30

150

3.5

0.4

55

18

30

150

0.4

55

18

30

150

4.5

0.4

55

18

30

150

0.4

55

18

35

150

0.4

60

18

35

150

3.5

0.4

60

18

35

150

0.4

60

18

35

150

4.5

0.4

60

18

10

35

150

0.4

60

18

11

40

150

0.4

65

18

12

40

150

3.5

0.4

65

18

13

40

150

0.4

65

18

14

40

150

4.5

0.4

65

18

15

40

150

0.4

65

18

16

45

150

0.4

70

18

17

45

150

3.5

0.4

70

18

18

45

150

0.4

70

18

19

45

150

4.5

0.4

70

18

20

45

150

0.4

70

18

The experimental program revealed several shortcomings of the experimental


setup. Some of these were due to the fact that the system was designed when the
thermodynamic cycle model was still in an early stage of development. Early modeling

73
relied on using the LKP equation of state, which necessitated several inaccurate
assumptions as discussed in Chapter 3.3 . Most significantly, the model assumed a perfect
desorption and separation process (i.e., the vapor leaving the separators consisted only of
pure CO2 in the saturated state). Since this was shown to have a significant impact on the
behavior and performance of the cycle, it would have been beneficial to have included
this analysis at the design stage. Furthermore, due to the lack of available fluid property
data, it was extremely difficult to size the heat exchangers. Ideally, a moving boundary
model would be used to determine the appropriate surface area requirements instead of
just a nominal capacity rating. The decision to proceed with the system construction was
driven by the desire of the project sponsor to obtain proof of concept results as soon as
possible.
In addition, several of the shortcomings were due to the necessity of using off-theshelf hardware and cost restrictions. The most significant operating limitations of the
system were as follows:

Pressure limitations of the equipment, particularly the heat exchangers,


limited the high side temperatures and concentrations that could be tested.

The heat exchanger pinch points, particularly for the desorber, were far
higher than predicted. In particular, the desorber was severely undersized,
so that the desorption process was not very effective. This meant the flow
rates run through the system had to be lowered significantly to get any
desorption of the CO2 at all.

The pump was severely oversized for this application due to low flow
rates being run.

74

The pump had extremely high fluctuations due to the plunger design,
which was possibly exacerbated due to the extremely low operational
speeds being run.

The expander seals failed extremely quickly due to the ingress of Acetone
into the expansion chamber. The same was true of the pump seals.

The expansion ratio fixed by the geometry of the expander was not ideal
for the given operating conditions. A design modification that included
suction and discharge valves to control the flow of gas into and out of the
expander is needed.

It was generally impossible to vary one input parameter during a test while
holding the remaining parameters constant. For example, it was extremely
difficult to vary the circulation ratio without changing the concentration of
the rich solution due to the ineffectiveness of the desorber.

The system was physically large compared to the power produced. This
resulted in significant heat losses and gains at locations other than the heat
exchangers.

Despite the limitations of the experimental system, it proved to be a valuable


research tool. The experimental test matrix was modified significantly to test a much
narrower range of operating conditions that the equipment was capable of handling. The
key results from the experimental testing are presented below.

75
5.2

Experimental Performance Trends

Table 5-2 summarizes the assumptions and device efficiencies when validating
the experimental test data against the thermodynamic model. Unless otherwise stated,
these assumptions were applied when generating the results described in this section.
Table 5-2: Key assumptions used in the thermodynamic model for experimental
comparison.
Description
Value
Temperature difference between heat sink and
1. Fixed according to
condenser/absorber outlet
experimental results
Temperature difference between heat source and
2. Fixed according to
evaporator/desorber outlet
experimental results
Heat sink temperature
3. Fixed according to
experimental results
Heat source temperature
4. Fixed according to
experimental results
Regenerator/internal heat exchanger effectiveness
5. 0.7
Pump isentropic efficiency
6. 0.25
Expander isentropic efficiency
7. 0.3
Negligible pressure drop in lines, separators, and heat
8.
exchangers
Negligible heat loss in lines, mixer, separator, pumps, and 9.
expander
Complete separation of liquid and gas phases
10.
Figure 5-1 compares the Second Law Efficiencies as a function of the Circulation
Ratio of the experimental results and the thermodynamic simulation model. Note that due
to the limitations in the experimental system and modeling tools discussed above, it was
necessary to match the conditions of each point in the simulation with those found while
running the experiments. For example, the pinch point temperatures between the source
and sink fluids and the working fluid outlet temperatures were prescribed due to the
inability to define an effectiveness parameter. Therefore, the absorber and desorber exit
temperatures (i.e., the lowest and highest cycle temperatures respectively) are identical in

76
the simulation and experimental results. The circulation ratio in the simulation model was
also set to match those observed in the experiments. These assumptions are acceptable
because the overall goal of the experimental program was to validate the accuracy of the
thermodynamic property evaluation from the Kunz and Wagner EOS, and by extension,
prove the validity of the thermodynamic model. Given these prescribed conditions, it can
be seen that there is good agreement between experimental and simulation results. In this
case, the data trend is in a direction where lower Circulation Ratios lead to higher Second
Law Efficiencies. This is consistent with Figure 3-8, which shows that there is an optimal
Circulation Ratio for a given rich solution concentration and available temperature
potential. In this case, the optimal Circulation Ratio is likely lower than 3.2. The trend
shown in
Figure 5-1, however, is weak. There is some scatter in the data, particularly
around a Circulation Ratio of 4, where lower Circulation Ratios do not necessarily lead to
higher Second Law Efficiencies. This is due to the inability to isolate parameters for
example, small differences in temperature for the same circulation ratio can have an
impact on the pressure ratio across the expander, which in turn can have a big impact on
overall cycle efficiency due to the equipment performance. Tight control of each
parameter requires optimized and reliable equipment, and was not generally possible for
the experimental study. Therefore, the overall goal of the experimental study was to
obtain a good match between experimental and simulation data for model validation,
rather than optimize the equipment and achieve good performance.

77

Figure 5-1: Comparison of theoretical and experimental Second Law efficiencies


as a function of the Circulation Ratio.
Figure 5-2 shows the desorber heat input (System Capacity) as a function of the
Circulation Ratio. In general, the expected trend would show that the Capacity decreases
as the Circulation Ratio increases. However, this is distorted in Figure 5-2 due to external
factors, most notably the fact that the rich solution concentration was not constant during
all the tests. This is due to the differences in charge inventories during the different
experimental runs. Different system operating points, and by extension, different system
capacities are possible at the same Circulation Ratio when the rich solution concentration
is not held constant (this may be inferred from Figure 3-7). There were also slight
variations in the source and sink temperatures which contribute to the distortion in the
trend. The three anomalous points that lie well outside the experimental uncertainty range
at low Circulation Ratios may be explained by poor desorber performance at these points,
i.e., for the given temperature, the desorber was not able to pull enough vapor out of the

78
solution to be consistent with model predictions. This is because the desorber was
severely undersized.

Figure 5-2: Comparison of theoretical and experimental Desorber Heat Input


(System Capacity) as a function of the Circulation Ratio.

Figure 5-3: Comparison of theoretical and experimental Second Law Efficiencies as a


function of the Desorber Exit Temperature.

79
Figure 5-3 shows the Second Law Efficiency as a function of the Desorber Exit
Temperature. Presumably the efficiency would increase as the exit temperature increases;
however this is clearly not the case in the experimental results. This is primarily due to
the fact that it was extremely difficult to isolate this variable. Higher temperatures tend to
lead to higher desorption rates, which in turn changes the Circulation Ratio and system
concentration levels. Therefore, the entire operating conditions of the equipment are
different, and the trend is significantly skewed. However, when the experimental
operating conditions are matched with the simulation model as in Figure 5-3, there is
good agreement between simulation and experimental data. This shows that the
thermodynamic basis of the model is robust.

Figure 5-4: Comparison of theoretical and experimental Expander Pressure Ratio as a


function of the Circulation Ratio.
Figure 5-4 shows the pressure ratio across the expander as a function of the
Circulation Ratio at the points for which reliable data were available. The trend

80
corresponds with that shown in Figure 3-15, which shows that higher Circulation Ratios
lead to higher Pressure Ratios. However, there are not enough data points that hold these
set of parameters in isolation to make a conclusive finding.

Figure 5-5: Comparison of theoretical and experimental Pump Work Input as a function
of the Circulation Ratio.

Figure 5-6: Comparison of theoretical and experimental Expander Work Output as a


function of the Circulation Ratio.

81
Figure 5-5 and Figure 5-6 show the pump and expander work as a function of the
Circulation Ratio. The required pump work would be expected to increase linearly with
the Circulation Ratio. However, as shown in Figure 5-2, this is distorted due to the rich
solution concentration not being constant and the slight variations in the source and sink
temperatures during the tests. The expander work would be expected to increase as the
Circulation Ratio decreases until an optimum ratio is reached, and in general, this trend is
followed. The distortions in the trend are largely due to the same pitfalls outlined above.
In both Figure 5-5 and Figure 5-6, however, there is broad agreement between
experimental and simulation data.

5.3

Experimental Program Summary

The experimental results for the ORCSC system provided a proof of concept
validation of the technology. The second law efficiencies of the cycle were generally
poor due to the numerous losses and limitations of the system. Isolation of parameters
was particularly difficult due to design shortcomings, the hardware limitations that are
inherent in early stage development work, as well as the number of variables that need to
be tightly controlled during system preparation and operation.
Nevertheless, an important finding was that there was largely good agreement
between the experimental and simulation data when all the boundary conditions were
matched. Generalizing this to broader measures of performance for the equipment,
however, remains a challenge. While the efficiency of the components and system as a
whole was not at a level needed for a viable commercial system, the results provided
confidence that the simulation model is robust.

82
The experimental study provided valuable insight into design and equipment
related aspects. For example, it was noted that the desorber was severely undersized
despite varying source temperatures, the desorber exit temperatures remained within a
narrow range. Furthermore, it was noted that a concentric tube-in-tube heat exchanger
design is not appropriate for facilitating a mass transfer process and a shell and tube
design would be far more suitable. Using a piston pump was found to cause large
fluctuations and vibrations in the system, despite the manufacturers assurances.
Teardown inspection of the rotating equipment revealed severe seal damage due to
contact with Acetone. This would discourage the use of Acetone as the absorbent in a
commercial system. The instrumentation was found to be adequate and performed
suitably during the experiments, with the exception of the torque cell, which was erratic
and required several data points to be discarded.
The experimental system also proved to be a useful test-bed for evaluating the
ORCSC concept. It pointed to key equipment that needs to be developed for the
technology to be successful, such as the need for an expander with a variable expansion
ratio to accommodate different operating conditions efficiently. It also provided physical
intuition for key parameters that are essential for controlling the system. A good example
is the Circulation Ratio, which is directly controlled be the relative speeds of the pump
and expander, and indirectly sets the concentrations of the system. Several of these
insights will be invaluable when developing ORCSC technology in the future.

83

CHAPTER 6. PERFORMANCE CHARACTERIZATION OF THE ORCSC SYSTEM

With the validity of the model having been confirmed by the experiments, the
model is used to characterize the performance of the ORCSC. This includes a sensitivity
analysis of various system variables on overall system performance. The experience from
the experiments is also used to determine cycle enhancements that would improve the
performance of the system. Therefore, this chapter is an extension of the work presented
in CHAPTER 3.

6.1

Parametric Analysis

The goal of the parametric analysis was to determine the sensitivity of the key
pieces of equipment, namely the pump, expander and heat exchangers, on overall cycle
performance. The performance metrics are the overall Second Law Efficiency and the net
power produced by the system for a given heat input of 15 kW (i.e., the desorber heat
input is 15 kW). Table 6-1 summarizes the baseline conditions used in the parametric
analysis. Unless otherwise specified, the conditions are held constant. Note that the
working fluid is Carbon Dioxide / Acetone at high concentrations ( rich = 0.5) .With these
baseline values, the system Second Law Efficiency is 11.93% with a net power output of
0.41 kW for the given 15 kW heat input.

84
Table 6-1: Baseline conditions for parametric analysis.
Baseline
Baseline
Parameter
Parameter
Value
Value
Tsource (C)

100

rich (-)

0.5

Tsin k (C)

15

CR (-)

TD , pinch (C)

10

turb (-)

0.8

TA, pinch (C)

10

pump (-)

0.6

QD (kW)

15

IHX

0.95

(-)

Figure 6-1 examines the impact on overall system Second Law Efficiencies as the
expander isentropic efficiency is varied from 45% to 90%, while the rest of the system is
fixed at the baseline conditions shown in Table 6-1. The overall system efficiency shows
the expected positive linear dependence on expander efficiency. Further, it is clear that
the overall system performance is highly dependent on the performance of the expansion
machine; therefore, small improvements in the expander efficiency result in a big
improvement in the overall Second Law Efficiency. For example, increasing the
expander efficiency from 70% to 80% results in an improvement in Second Law
Efficiencies from roughly 8.5% to 12%, a performance improvement of approximately
40%. This shows the importance of designing an optimized expander for the ORCSC
system.

85

Figure 6-1: Net power output and Second Law Efficiency as a function of expander
isentropic efficiency.
Figure 6-2 examines the effect of varying the pump isentropic efficiency from 40%
to 90% on the overall system Second Law Efficiency. Analogous to the expander
efficiency, the overall system efficiency shows a positive linear dependence on the pump
efficiency. Like the expander, the overall system performance is highly dependent on the
performance of the pump. For example, increasing the pump efficiency from 50% to 60%
results in an improvement in Second Law Efficiencies from roughly 9% to 12%, a
performance improvement of roughly 33%. Unlike the expander, however, there is
perhaps less scope to improve the performance of the pump due to technical hurdles and
design limitations imposed by the requirement of high flow rates as well as high pressure
lifts required from the pump in an ORCSC system.

86

Figure 6-2: Net power output and Second Law Efficiency as a function of pump
efficiency.
Figure 6-3 shows the effect of varying the desorber pinch point temperature,
defined in Equation ( 3-17 ) as the difference between the heat source inlet temperature
and the desorber exit temperature (maximum working fluid temperature), on the system
Second Law Efficiency. Since the rich solution concentration and source temperature is
fixed (the intrinsic assumption being that the heat source is a constant temperature
reservoir), the desorber pressures vary along with the pinch point temperatures. Note that
it is assumed that the desorber is able to facilitate the desorption process, i.e., pull the
absorbate vapor out of the absorbent solution, regardless of the pinch point temperature.
Therefore, Figure 6-3 illustrates a measure of the heat transfer effectiveness of the
desorber rather than a more complete heat and mass transfer performance parameter.
Nevertheless, Figure 6-3 shows that the pinch point temperature can have a significant
impact on the overall cycle performance. Increasing the pinch point temperature from

87
5C to 15C degrades the system performance from roughly 12.5% to 11%, a drop of
12%.

Figure 6-3: Net power output and Second Law Efficiency as a function of the desorber
pinch point temperature.
Figure 6-4 illustrates the effect of varying the absorber pinch point temperature,
defined in Equation ( 3-18 ) as the difference between the heat sink inlet temperature and
the absorber outlet temperature (minimum working fluid temperature), on the system
Second Law Efficiency. Analogous to the desorber pinch point temperature, the overall
system efficiency degrades as the absorber pinch point temperature increases. For
example, increasing pinch point temperature from 5C to 15C degrades the system
performance from roughly 14.7% to 9.35%, a drop of 36%. This shows that the overall
system performance is extremely sensitive to the absorber performance significantly
more so than the desorber performance. However, absorption is easier to facilitate from a
mass transfer standpoint, particularly with a working fluid like Carbon Dioxide, because

88
the high pressures force the absorbate into the solution. Nevertheless, designing an
appropriate absorber for the ORCSC system is critical.

Figure 6-4: Net power output and Second Law Efficiency as a function of the absorber
pinch point temperature.
Figure 6-5 illustrates the effect of varying the internal heat exchanger
effectiveness, defined in Equation ( 3-19 ), on the system Second Law Efficiency. Since
the internal heat exchanger only facilitates heat exchange between two streams in the
liquid phase, it is possible to define a heat transfer effectiveness in the conventional
manner. Clearly, the overall system efficiency increases as the effectiveness increases;
however, the impact is relatively small. For example, increasing the heat exchanger
effectiveness from 70% to 80% improves the Second Law Efficiency from roughly 9.85%
to 10.6%, an increase of 7%.

89

Figure 6-5: Net power output and Second Law Efficiency as a function of the internal
heat exchanger effectiveness.
The parametric studies clearly show that the ORCSC is intrinsically dependent on
component efficiencies. The performance of the rotating equipment, in particular, was
shown to have a substantial impact on the performance of the ORCSC system as a whole.
The performance of the absorber and desorber is difficult to characterize due to the mass
transfer process in addition to the heat transfer. However, by utilizing pinch point
temperature as a performance parameter, it was shown that heat exchanger performance
is also important to the overall system performance. The absorber performance is perhaps
the most critical of the three heat exchangers; however, absorption is easier to facilitate
than desorption because the high system pressures in the ORCSC force the absorbate into
the solution. This is consistent with experimental observations, which showed
significantly lower pinch point temperatures in the absorber compared to the desorber.
Translating pinch point temperatures into broader design metrics such as required surface
area, however, remains a challenge.

90

CHAPTER 7. CONCLUSIONS AND FUTURE WORK

7.1

Conclusion

A novel Organic Rankine Cycle with Solution Circuit using Carbon Dioxide /
Acetone as the working pair was studied in detail. While the cycle configuration,
variously called an Absorption-Rankine cycle or Kalina cycle in the literature, is not
entirely new, the choice of working fluid pair marks a departure from traditional fluids
such as Ammonia / Water. This expands the potential applications of the technology, and
offers important solutions required for the adoption of the technology.
A thermodynamic analysis of the ORCSC was performed using the Kunz and
Wagner equation of state programmed in Refprop 9.0 and optimized with an EES-based
computer model. It was found that the ORCSC using a Carbon Dioxide / Acetone
working pair offers no significant efficiency improvements over a conventional Organic
Rankine Cycle (ORC) using only Carbon Dioxide as the working fluid. Furthermore, the
ORCSC with a Carbon Dioxide / Acetone working pair has significantly lower
performance than an ORC using conventional working fluids such as pentane or R245fa.
The local peak observed at extremely low concentrations of Carbon Dioxide was higher
than the peak observed in the traditional interpretation of an ORCSC with high Carbon
Dioxide concentrations; however, it does not perform as well as a traditional ORC with
pure Acetone to which the low concentration ORCSC is comparable. Nevertheless, the

91
use of a zeotropic mixture in the ORCSC provides the ability to use the temperature glide
to match the temperature profiles of the source and sink fluids, and leads to significantly
lower working pressures than using the pure absorbate in an ORC. The zeotropic mixture
also allows for capacity control simply by changing the concentration of the working
fluid mixture. This may lead to higher overall system efficiencies when coupled with
sources that have varying heat input temperatures or loads, such as waste heat streams
from diesel engines, power plants or other industrial processes.
An experimental test stand was constructed in order to validate the simulation
model. Data from several test runs showed that the efficiency of the load stand was
generally poor, and many other performance metrics showed erratic behavior. This was
largely due to design shortcomings and the use of readily available off-the-shelf
components that were not optimized for the ORCSC system. Isolation of parameters was
particularly difficult given the number of variables that need to be tightly controlled
during system operation. Nevertheless, there was generally good agreement between the
experimental and simulation data when all the boundary conditions were matched.
Therefore, the experimental results provided a proof-of-concept validation of the
technology.
The validated model was extended to include a parametric study of various
system variables on overall system performance. It showed that the ORCSC is dependent
on component efficiencies, and that the performance of the expander and pump are
particularly critical to the performance of the ORCSC system. Amongst the heat
exchangers, the performance of the absorber had the greatest impact on the overall
system performance.

92
It is clear from this study that a range of practical considerations need to be taken
into account and weighed together with the thermodynamic performance when evaluating
the feasibility of ORCSC technology. The ORCSC offers some potential practical
advantages, such as capacity control and lower working pressures, which may outweigh
the added cost and complexity of the system in certain applications. However, there are
several performance and practical tradeoffs when comparing ORCSC and ORC systems,
and any choice between the two is likely to be dictated by the application. The maturity
of ORCSC technology and associated body of literature is limited, and further work
needs to be pursued in this area before widespread adoption of the technology is possible.

7.2

Recommendations for Future Work

While it is clear that the ORCSC offers important solutions needed for widespread
adoption of low-grade heat recovery technology, there is little information available in
the literature that compares different working fluids in this cycle configuration. This is a
critical area that needs to be addressed, and an appropriate starting point would be to
examine the working pairs used in absorption technology. A fundamental difference is
that absorption technology generally favors working pairs with far lower generation
(desorption) temperatures than would be appropriate for ORCSC technology. Finding a
variety of appropriate working pairs for the ORCSC system would greatly expand the
field of application, and hasten adoption of the technology.

93
Of the three major mechanical components used in the ORCSC, heat exchangers
and pumps have been widely studied in the literature, and several commercial products
exist in the market. However, expansion machines used in ORC and ORCSC technology
have not been widely studied, and few commercial designs are available. The parametric
study performed as part of this work showed that the overall system performance is
highly dependent on the performance of the expansion turbine. Therefore, a concerted
effort is needed to develop an expander design for ORCSC technology. This would likely
require an analytical model of the expansion machine. The model can leverage previous
work performed by Lemort et al. (2008), Mathison et al. (2008), and Bradshaw et al.
(2011), in developing analytical and semi-empirical positive displacement machinery
models. The analytical model of the expander would allow for the optimization of the
expander design, and subsequently to the optimization of the ORCSC system.
Another future effort required for the ORCSC system is to find robust ways in
which to control the system without frequent human input. This requires identification of
key parameters that are controllable by mechanical equipment, as well as the design and
development of the equipment itself. For example, any application that involves the
ORCSC to exercise capacity control would require the ability to measure the
concentration of the working fluid in-situ. There are no commercial products that are
explicitly designed to do this, although the development of void fraction technology for
this purpose is promising. There are few studies in the literature that explicitly examine
the control of ORC or ORCSC systems, and there is a large scope for advancement in this
area.

94
Specific to the work performed as part of this project, a logical next step is to
expand the simulation model. While a thermodynamic model has been developed for the
ORCSC system, a physical model that includes performance parameters of the
mechanical components needs to be developed. This includes developing either
mechanistic or map-based models for the rotating equipment and moving boundary
models for the heat exchangers. Prior to developing moving boundary models for the heat
exchangers, however, performance metrics that are capable of capturing the heat and
mass transfer processes need to be identified and analyzed. The physical model could
then be validated using experimental data obtained from the existing test setup at the Ray
W. Herrick Laboratories. A physical model would provide the opportunity to conduct a
fuller assessment of system performance, and the model could be modified to match the
equipment design as well as environmental parameters and operating conditions. Such an
investigation would develop an enhanced understanding of the overall system sensitivity
to design variables.

LIST OF REFERENCES

95

LIST OF REFERENCES

1. Altenkirch, E., 1950, Kompressionskltemaschine mit Lsungskreislauf,


Kltetechnik, 2 (1950) 10, pp. 251-259, 2 (1950) 11, pp. 279-284, and 2 (1950)
12, pp. 310-315.
2. ASHRAE, 2005, Datacom Equipment Power Trends and Cooling Applications,
ASHRAE, Editor.
3. Badr, O., OCallaghan, P. W., & Probert, S. D., 1990, Rankine-Cycle Systems
for Harnessing Power from Low-Grade Energy Sources, Applied Energy, 36(4),
263-292. doi:16/0306-2619(90)90002-U.
4. Bradshaw, C. R., Groll, E. A., and Garimella, S. V., 2011, A Comprehensive
Model of a Miniature-Scale Linear Compressor for Electronics Cooling,
International Journal of Refrigeration, vol. 34, pp. 6373.
5. Davidson, T. A., 1977, Design and Analysis of a 1 kW Rankine Power Cycle,
Employing a Multi-Vane Expander, for use with a Low Temperature Solar
Collector, Massachusetts Institute of Technology.
6. Enertime, 2011, The Organic Rankine Cycle and its Applications, Retrieved
from http://www.cycle-organique-rankine.com.
7. Goswami, D., Y.; Xu, F.; Bhagwat., S., 2000, A Combined Power/Cooling
Cycle, Energy 25 (2000), pp. 233246.
8. Groll, E A, Radermacher, R, 1994, Vapor Compression Cycle with Solution
Circuit and Desorber/Absorber Heat Exchange, ASHRAE Trans, 100 (1): 73e83.
9. Hung, T.-C., 2001, Waste Heat Recovery of Organic Rankine Cycle Using Dry
Fluids, Energy Conversion and Management, 42(5), 539-553. doi:16/S01968904(00)00081-9.
10. Hugenroth, J., 2006, Liquid Flooded Ericsson Cycle Cooler, Ph.D. Thesis, Purdue
University.

96
11. Incropera, F. P., DeWitt, D. P., Fundamentals of Heat and Mass Transfer, 5th
Edition, John Wiley & Sons, ISBN-13: 978-0471794714.
12. Kalina A. I., 1984, ASME paper 84-GT-173, Fairfield, NJ.
13. Klein, S, 2012, Engineering Equation Solver, F-Chart Software.
14. Krishna A, Groll E A, Garimella S V., 2011, Organic Rankine Cycle with
Solution Circuit for Low-Grade Waste Heat Recovery, Alternative Sources /
International Sorption Heat Pump Conference, April 5-7, Padua, Italy, I54.
15. Kunz, O., Klimeck, R., Wagner, W., Jaeschke, M., 2007, The GERG-2004
Wide-Range Equation of State for Natural Gases and Other Mixtures, GERG
Technical Monograph 15. Fortschr.-Ber. VDI, VDI-Verlag, Dsseldorf.
16. Kyle, B. G., Chemical and Process Thermodynamics, 3rd Edition, Prentice Hall
PTR, ISBN-13: 978-0130874115
17. Lemmon, E W, Huber, M L, McLinden, M O, 2010, NIST Standard Reference
Database 23: Reference Fluid Thermodynamic and Transport PropertiesREFPROP, Version 9.0, National Institute of Standards and Technology,
Standard Reference Data Program, Gaithersburg, 2010.
18. Lemort, V., Bell, I., Groll, E., Braun, J., 2008, Analysis of Liquid-Flooded
Expansion Using a Scroll Expander, Proceedings of the 19th International
Compressor Engineering Conference at Purdue University.
19. Little, A., B.; Garimella, S., 2009, Comparative Assessment of Alternative
Cycles for Waste Heat Recovery and Upgrade, Proceedings of the ESC 2009:
ASME 3rd International Conference on Energy and Sustainability, July 19-23, San
Francisco, CA.
20. Maloney, J. D. and Robertson. R.C., 1953, Thermodynamic Study of Ammonia Water Heat Power Cycles, Oak Ridge National Laboratory, Oak Ridge,
Tennessee.
21. Mathison, M.M., Braun, J.E., Groll, E.A., 2008, Modeling of a Two Stage
Rotary Compressor, HVAC&R Res. 14 (5), 719e748.

97
22. Monahan, J., 1976, Development of a 1-kW, Organic Rankine Cycle Power
Plant for Remote Applications, Presented at the Intersociety Energy Conversion
Engineering Conference, New York.
23. Mulroy, W. J.; Domanski, P. A.; Didion, D. A., 1993, Glide Matching with
Binary and Ternary Zeotropic Refrigerant Mixtures, Part 1. An Experimental
Study, Thermal Machinery Group, National Institute of Standards and
Technology, Gaithersburg, MD 20899, USA.
24. Plocker, U., Knapp, H. and Prausnitz, J., 1978, Calculation of High Pressure
Vapor-Liquid Equilibria from a Corresponding States Correlation with Emphasis
on Asymmetric Mixtures, Ind. Eng. Chem. Process. Des. Dev., 17(3): 324-332.
25. Probert, S. D., Hussein, M., OCallaghan, P. W., & Bala, E., 1983, Design
Optimization of a Solar-Energy Harnessing System for Simulating an Irrigation
Pump, Applied Energy, 15(4), 299-321. doi:16/0306-2619(83)90059-4.
26. Quoilin, S., 2011, Sustainable Energy Conversion Through the Use of Organic
Rankine Cycles for Waste Heat Recovery and Solar Applications, Ph.D. Thesis,
University of Lige.
27. Robbins, T., Garimella, S., 2010, Low-Grade Waste Heat Recovery for Power
Production using an Absorption-Rankine Cycle, Proceedings of International
Refrigeration and Air Conditioning Conference, ASHRAE/IIR: No. 2517.
28. WEC, 2006, World Energy in 2006, World Energy Council, London, UK.
29. Woodland B J, Braun J E, Groll E A, Horton W T, 2010, Performance Benefits
for Organic Rankine Cycles with Flooded Expansion and Internal Regeneration,
Proceedings of the 2010 International Refrigeration and Air Conditioning
Conference, Purdue University, R022.
30. Zheng, D.; Chen, B.; Qi, Y.; Jin, H., 2006, Thermodynamic Analysis of a Novel
Absorption Power/Cooling Combined-Cycle, Applied Energy 83 (2006), pp.
311-323.

APPENDICES

98
Appendix A

Instrument Calibration Procedures

Since all measurement devices including thermocouples, pressure transducers,


flow meters and torque cells were harvested from old test stands, calibration of the
measurement devices was necessary in order to acquire accurate measurement data
during the experimental tests. The objective of the calibration procedure was to obtain a
plot relating certain reference values against the signal output by the measurement device.
Using the data points, a linear regression was used to determine the calibration equation
which linked the signal outputs with the measured value.

Temperature Calibration
As noted previously, temperatures at each state point in the cycle were measured
using OMEGA T-type thermocouples. T-type thermocouples consist of two different
materials, copper and constantan, that generate a voltage corresponding to a certain
temperature. In order to obtain a calibration curve, signal outputs from two fixed points
were measured and a linear regression was applied to the data. In this case, the two
chosen points were the freezing temperature (T = 0 C) and boiling temperature (T =

99.75 C at 1.005 bar ambient pressure) of distilled water. The freezing temperature was

set using an ice bath containing distilled water ice cubes in equilibrium with distilled
liquid water. The thermocouples were then immersed into the ice bath and the voltage
signal was recorded. Similarly, in order to obtain the second calibration point, a water
heater was used to boil the distilled water, and the signal was recorded from the

99
immersed thermocouple. Figure A 1: Sample calibration data for a thermocouple. Figure
A 1 shows a sample calibration curve for a thermocouple.

120

Temperature (C)

100
80
60

Calibration Equation:
T(C) = 1.0047V(mV) - 0.253
R = 1

40
20
0

20

40

60

80

100

120

Voltage (mV)
Figure A 1: Sample calibration data for a thermocouple.
Pressure Calibration
The pressure transducers were calibrated against a reference pressure transducer
using an apparatus available at the Herrick Laboratories. A nitrogen tank provided the
required pressure, and the apparatus contains an accurate pressure reading device
(OMEGA PCL 1B) against which the pressure transducer can easily be calibrated. As a
controlled pressure was applied using valving connected to the nitrogen tank, an average
voltage output was taken via LabVIEW, and the reference pressure was read from the
calibration device. Subsequently, the pressure was raised in roughly equal increments
until the maximum desired pressure (650 psi) was reached. When the maximum pressure

100
was reached, the pressure was lowered in equal increments in order to detect any
hysteresis in the measurements. Average values were computed for the output values
from the increasing and decreasing pressure measurements. For these average values, a
trend line was computed that serves as the calibration equation. Figure A 2 shows a
sample calibration curve for a pressure transducer. Note that for one of the pressure
transducers that was tested, significant hysteresis effects could be observed.
Consequently, this device could not be used. While reading the voltage outputs, it was
also observed that this defective pressure transducer also showed much higher noise in
the voltage output signals when compared to the other pressure transducers.
It was also observed that grounding was particularly important when using
pressure transducers with a voltage range of 0 10 mV. These output voltages were
easily distorted by other electrical devices when compared to pressure transducers having
output voltages ranging from 1 6 V. It is important to note that the pressure transducers
require an excitation voltage to function. This excitation voltage may not be changed
after calibrating the pressure transducer. This is because the output voltages from the
pressure transducer vary with different excitation voltages, even if the applied pressure is
constant. It is also important to note that the pressure calibration exercise was conducted
at room temperature. Prior experience has demonstrated that voltage outputs vary slightly
depending on the temperature, which may lead to slight error in the pressure readings.

101
700

Calibration
Equation:

Reference Pressure (psi)

600

p = 83192V + 5.2842
R = 1

500
400

increasing
pressure
decreasing
pressure
average

300
200
100
0

Linear
(average)
0

0.002

0.004
0.006
0.008
Voltage output (V)

0.01

Figure A 2: Sample calibration data for a pressure transducer.


Flow Meter Calibration
The mass flow meters were calibrated using utility water supply available at the
Herrick Laboratories. Although the mass flow meters were to be used to measure a
Carbon Dioxide / Acetone mixture, the calibration is valid since the flow meters also
measure the density of the fluid during operation. Furthermore, since Micro Motion
guarantees the calibration of newly purchased flow meters and provides calibration
curves in their data sheets, a recently purchased flow meter was inserted in series with the
flow meter to be calibrated in order to validate the measured flow rates.
The mass flows were measured by running the utility water through the flow
meters and into a bucket for a certain time interval. The mass of the water in the bucket
was then measured using a high accuracy scale. Dividing the mass of the water with the
measured time interval gives the mass flow rate. While filling the bucket and recording
the time interval, a LabVIEW program recorded the signal outputs from the flow meter.

102
In parallel, the values on the Micro Motion transmitter display were also noted. Figure A
3 shows a sample calibration curve for a flow meter. Note that FM733309 was the
reference flow meter recently purchased and calibrated by Micro Motion, and that FM
2203437 was the flow meter being calibrated for the experimental breadboard system.
Although the signal outputs are significantly different, comparing the data acquired by
the two flow meters in series show that the flow rates differ by 0.0015% on average.

Measured mass flow rate (kg/s)

0.35

Calibration Equation for


FM 2203437:

0.3

m = 23.748A - 0.0928
R = 0.9996

0.25
0.2

FM2203437

0.15

FM733309

0.1

display

0.05
0

Linear
(FM2203437)
0

0.005
0.01
0.015
Current output (A)

0.02

Figure A 3: Sample calibration data for a mass flow meter.


Torque Cell Calibration and Outputs by the Drives
In order to measure the work output from the expander, a rotating torque sensor
was placed in between the turbine shaft and belt drive as shown in Figure A 4. The
objective of the torque cell calibration was to obtain a plot relating the applied moment
against the voltage output. The output values from the torque cell range from 0 2 mV.

103
Therefore, a Strain Gage Signal Conditioner amplifies the output signal before the output
is read by the DAQ card.

Figure A 4: Top view of expander and torque cell.


The calibration exercise involved applying a torque on the torque cell using an
aluminum bar with a coupling on one side of the bar (see Figure A 5). A level was used
to fix the bar in a horizontal position and calculate the correct radius for the center of
mass. The center of mass was calculated to be 0.2948 meters (rcenter) away from the axis
of the shaft. With this information and the known weight of the bar, the applied torque
was calculated as follows:
Tor = mtotal * g * rcenter

( AA-1 )

where mtotal represents the mass of the bar and the added weight, and g the gravitational
acceleration. In subsequent steps, additional weight was added to the center of mass. This
was accomplished by fixing a string to the center of mass, and adding weights in 400
gram increments. Since the string is located at the same center of mass as the bar, the
same radius could be used, and the moment generated by the weights could easily be
added to the one applied by the bar. While taking measurements, one shaft was restrained

104
and kept in a fixed position while the other was allowed to turn in order to measure the
torque using the strain gages inside the torque cell. Figure A 6 shows a sample calibration
curve for the torque cell.

Figure A 5: Calibration of the torque cell.

Measured Torque (Nm)

14
12

Calibration Curve:

10

T = 6.2475V - 2.4903
R = 0.9999

torque

6
4

Linear (torque)

2
0

0.5
1
1.5
2
Amplified Voltage Output (V)

2.5

Figure A 6: Sample calibration data for torque cell.

105
Appendix B

Wiring Diagram for Data Acquisition System

106
Appendix C

1.
2.
3.

4.

5.
6.

7.

Procedures for Charging and Discharging the ORCSC System

Charging of System
Fill the charging cylinder with the predetermined mass of acetone.
Pull a vacuum on the system by connecting a vacuum pump to the plug
valve located below the separators (V11). Once the system is under
vacuum, close the plug valve (V11).
A siphoning effect will be used to charge the acetone. Feed the
charging tube into the acetone container until the tube is full of acetone.
The valve located at the end of the charging tube may be used to drain
any air bubbles.
Once the tube is free of any air bubbles, place the acetone container on a
weighing scale and record the weight. Then connect the charging tube to
the charging valve (V15) and allow the vacuum to pull in the required
amount of acetone.
Once the system has been charged with the required amount of acetone,
the charging valve (V15) and the valve on the charging tube may be
closed, and the charging tube disconnected.
The system is then loaded with a prescribed amount of CO2 by
connecting the CO2 cylinder to the charging valve located on top of the
receiver (V15). The CO2 cylinder may need to be opened slightly in
order to flush the connection tube with CO2 before charging the system
with CO2. Alternatively, a vacuum pump may need to be used to create a
vacuum on the connection lines.
Once the lines are under vacuum, close the valve to the pump. Slowly
open the plug valve (V15) and the valve on the CO2 bottle. CO2 will
enter the system due to the pressure difference between the CO2 bottle
(high pressure) and the system (low pressure). Successive mass
measurements of the bottle using an appropriate weighing scale may be
required in order to determine if the system has been charged with the
required amount of CO2. If needed, actively cool the system by running
the cooling water loop in the absorber to draw more CO2 from the bottle
into the system.

Discharging of System (The system needs to be discharged for long off-cycle


periods)
1. Ensure system is shutdown according to sections E and F in the ORCSC
startup and operating guide.

107
2. Slowly open valve atop the separator (V7). The CO2 in the system will
be purged through the vent line. Note that some acetone fog may
accompany the CO2. Thus, use a bucket below the vent line to capture
any acetone. If heavy droplets are observed, the valve (V7) is open too
far and needs to be closed accordingly. This process is performed until
all the CO2 is drained from the system and there is pressure equilibrium
between the system and surroundings.
3. Once the CO2 is drained from the system, the valves located at the lowest
reaches of the setup may be used to gravity-drain the acetone into a
containment vessel. These are V4, V11 and V16.

108
Appendix D

Procedures for Operating the ORCSC System

Please refer to Figure 4-1 and Figure 4-2 for locations of the control equipment
referenced in the procedures below.
1.
2.
3.
4.
5.

Pre-check
Secure area with chain and place appropriate warning signs.
Ensure area is clean and free of any trip hazards.
Complete a test rig walk down and check for any pooling of liquids.
Confirm operation of acetone alarms.
Compare current system shut-off pressure to last recorded shut-off
pressure to evaluate if leaks have occurred.

Pre-Startup
1. Check that the 240 V transformer breaker is off. This is located on the
wall across from test stand, between the two psychometric chambers, and
is labeled Organic Rankine Cycle.
2. Check that the table breaker is off. This is the large red lever-switch
underneath left front side of table.
3. Check that the Data Acquisition and Safety Systems are off and
unplugged from any power sources.
4. Check that the steam (ST1, ST2, ST3 and ST4) and water valves (W1)
are closed.
5. Check that the pump (V5) and expander bypass (V14) valves are open.
6. Close pump (V1, V2) and expander (V8, V9) shutoff valves. These have
been left open to equalize the pressure in the system.
7. Ensure expansion valves (V12, V13) are fully open.
Startup
1. Attach steam and water hoses and place into appropriate drains.
Measurement equipment:
2. Plug in the data acquisition system. Switch on the data acquisition
system (DAQ).
3. Switch on the Agilent DMM. Start LabVIEW and run the VI.
4. Check pressure readings via LabVIEW; ensure nominal values.
Switches:
5. Plug in the main power line to the test stand and then switch on the
transformer breaker (located between the psychometric chambers).

109
6. Switch on table breaker.
Prepare the Absorber Cycle:
7. Open the shutoff valves (W1) on the water line. Check for leaks.
8. Ensure that there are no kinks in cold water hose.
Prepare the Desorber Cycle / Glycol Loop
9. Start Glycol pump by plugging in the power cord
10. Open Steam / Hot Water shutoff valves (ST1, ST2, ST3 and ST4). Check
for Leaks
11. Adjust steam pressure to desired PSI via regulator on wall behind table.
Adjust the steam regulator to ensure there is enough vapor in the
separator.
12. Adjust Glycol temperature by a combination of adjustments to the 3-way
valve (T1) and steam valves (ST2, ST3, ST4).
Prepare the ORCSC Cycle:
13. Push Stop button on both motors if they show rpm greater than zero.
14. Open shutoff valves (V1,V2) on the suction and discharge sides of the
pump.
15. Start pump at 100 rpm via motor controller keypad.
16. Check sight glass at the bottom of the receiver and ensure full liquid flow
into the pump (not two-phase). If the flow is two-phase for more than 20
sec., stop the pump. Increase cooling water flow to subcool working
fluid at pump inlet.
17. Achieve desired mass flow rate by slowly closing the bypass valve (V5)
and increasing pump rpm.
18. Adjust mass flow rate of glycol to match desired heat flux.
19. The expansion valves (V12, V13) may be adjusted, if necessary, to
achieve a pressure equilibrium pressure between the weak solution
stream and the expander bypass stream. Ensure that working fluid
pressures reach stable values.
20. Check that the working fluid is boiling and condensing properly by
repeated checks of the sight glasses on the separator and receiver. The
steam and water mass flow rates may have to be adjusted to achieve this.
21. Open the shutoff valve (V9) on the expander outlet slowly.
22. Open the shutoff valve (V8) on the expander inlet slowly.
23. Start expander at 500 rpm and increase incrementally to approximately
2000 rpm.
24. The expansion valves (V12,V13) may be adjusted to ensure pressure
equilibrium between the weak solution stream and the expander CO2
stream.
25. Watch the working fluid pressures graph on LabVIEW while slowly
closing the expander bypass valve (V14). Be careful that the pressures do
not rise near the overpressure limit (50 bar gauge). The expansion valves
(V12, V13) should once again be adjusted as necessary.

110
26. Check again for two-phase flow into the pump and turbine. Periodically
check this condition, and shut down the system if necessary.
27. Steam flow may need to be adjusted once expander is online in order to
maintain the pressure difference between high and low sides.
Running/taking data
Check temperatures, pressures and flows:
1. Monitor system pressures, mass flow rates, motor torques, and steam
temperatures for unexpected fluctuations. Shut down the system if these
occur.
Adjusting system variables to take data points:
2. Vary the pump and expander rpm to the desired values according to the
test matrix. Ensure that pressure limits are not exceeded.
The expander has a range up to 5000 rpm.
The pump has a range up to 2000 rpm.
Shut Down
3. Open the expander bypass (V14). Spin down expander to 1000 rpm
using keypad arrows. Press Stop on expander motor controller.
4. Close steam valves (ST1, ST2, ST3 and ST4).
5. Allow pump to briefly cool heat exchanger (until a temperature of 75 C
is achieved).
6. Spin down pump to 100 rpm using keypad arrows. Press Stop on
pump motor controller.
7. Press the Stop button on the safety system.
8. Bleed steam line and allow the glycol loop to run in order to cool down
the system.
9. Once the glycol loop is cool enough, the glycol pump may be turned off
by unplugging the power cord.
10. Close the shutoff valve on the water line (W1).
11. Shut down power to table via table and transformer breakers.
12. Disconnect Table power cord and steam and water hoses.
Cool Down
1. Leave chain up until cycle temperatures are around 50 C.
2. Monitor temperatures of Rankine Cycle with LabVIEW until all
temperatures are below 50 C.

111
Emergency Shut Down
Hit Emergency Stop Button. Motors will coast to a stop. The expander will
likely over rev to release system pressure since the motor now applies no
load. Open bypass valve if possible to stop expander (V14). Kill power to
the system via the table breaker (Breaker 1) and the Organic Rankine
Cycle breaker. Turn off steam input at the source if safe (ST 1). In case of
a major leak, purge the charge using the drain line.

112
Appendix E

Run
index
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

T1
(C)
85.5
85.5
85.4
85.4
87.7
87.7
95.2
95.3
95.3
87.5
87.5
87.5
87.6
87.6
85.8
85.8
85.8
85.8
85.8
85.8

Data from Experimental Testing

Table E-1: Experimental temperature data.


T2
T3
T4
T5
T6
T7
(C)
(C)
(C)
(C)
(C)
(C)
71.0
23.7
26.8
59.5
24.2
18.3
71.0
23.8
27.0
59.4
24.3
18.2
70.8
23.9
27.4
59.1
24.3
18.1
70.8
23.9
27.5
59.3
23.9
18.1
73.3
24.8
28.0
60.4
25.2
19.2
73.3
24.8
28.2
60.4
25.1
19.1
78.2
24.3
29.9
64.6
24.2
19.8
78.3
24.3
30.0
65.1
24.3
20.1
78.3
24.3
30.0
64.7
24.3
20.0
75.2
29.5
28.2
66.8
24.8
17.7
75.2
28.9
28.1
66.8
24.6
17.7
75.3
28.5
28.1
67.1
24.5
17.6
75.2
28.6
28.1
66.9
24.8
17.6
75.2
28.6
28.1
67.0
24.7
17.5
73.1
23.4
27.8
58.9
24.9
16.7
73.0
23.4
27.8
58.9
25.3
16.5
73.0
23.4
27.8
59.0
25.5
16.5
73.0
23.7
27.7
58.9
25.2
16.4
73.0
23.7
27.7
58.9
25.2
16.4
73.0
23.7
27.7
58.8
26.0
16.5

T8
(C)
58.3
57.4
57.9
58.3
58.9
59.6
62.6
63.3
62.7
64.9
64.9
66.1
63.9
65.7
57.0
56.9
57.9
56.9
57.1
56.9

T9
(C)
26.0
26.1
26.6
26.7
26.8
27.0
28.9
28.9
28.9
27.3
27.3
27.2
27.4
27.4
26.6
26.6
26.6
26.5
26.6
26.5

113

Run
index
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

Table E-2: Experimental temperature data continued.


T10
T11
T12
T13
T14
T15
(C)
(C)
(C)
(C)
(C)
(C)
42.4
55.4
23.5
35.9
40.5
58.1
42.2
55.2
23.6
35.9
40.3
58.0
42.5
54.8
23.5
35.4
39.6
57.7
42.9
54.7
23.5
35.6
39.7
57.7
44.1
56.1
24.2
37.2
41.6
58.7
44.0
55.9
24.2
37.2
41.3
58.7
46.8
59.5
23.7
39.1
43.3
63.1
47.0
59.9
23.6
39.2
43.6
63.3
46.7
59.7
23.5
39.2
43.5
63.3
46.8
62.8
28.6
38.2
43.7
64.6
46.7
62.6
27.6
38.2
43.7
64.5
46.5
62.6
27.1
38.0
43.6
64.5
46.5
62.7
27.2
38.2
43.7
64.5
46.4
62.8
27.3
38.2
43.7
64.5
40.7
48.8
23.2
36.3
42.9
57.8
40.1
49.2
23.4
36.6
43.1
57.6
40.5
49.2
23.4
36.5
43.1
57.6
40.8
49.2
23.3
36.5
43.0
57.8
40.4
49.1
23.4
36.5
43.1
57.7
39.9
49.1
23.3
36.4
43.0
57.5

T16
(C)
14.4
14.4
14.4
14.4
15.5
15.5
15.3
15.4
15.3
14.9
14.9
14.9
15.0
14.9
12.4
12.2
12.2
12.2
12.2
12.1

T17
(C)
24.5
24.7
24.7
24.9
25.5
25.5
27.0
27.0
26.9
25.2
25.1
25.4
25.9
25.9
23.3
23.4
24.2
23.4
24.0
23.9

114
Table E-3: Experimental pressure data.
Run
P1
P2
P3
P4
P5
index (kPa)
(kPa)
(kPa)
(kPa)
(kPa)
1
3321.2 3357.5 2866.7 2990.0 3421.6
2
3315.1 3411.9 2858.1 2967.3 3411.9
3
3322.5 3436.7 2843.3 2953.1 3422.5
4
3354.2 3402.0 2854.0 3007.1 3448.8
5
3383.0 3503.9 2903.8 3086.9 3488.4
6
3368.6 3514.8 2850.2 3021.1 3474.9
7
3719.4 3868.6 3063.8 3275.5 3838.5
8
3694.8 3754.1 3041.9 3231.6 3820.6
9
3695.6 3748.1 3036.9 3216.4 3821.2
10
3457.3 3508.9 3153.2 3295.7 3576.1
11
3450.5 3554.1 3160.2 3317.2 3570.3
12
3466.4 3514.4 3175.8 3315.6 3585.6
13
3460.6 3551.4 3186.1 3328.2 3583.9
14
3466.4 3533.5 3213.9 3350.3 3590.4
15
3926.5 3899.6 2900.9 2950.4 4045.2
16
3920.3 3804.1 2908.1 2961.6 4030.3
17
3913.0 3993.3 2911.4 2961.7 4051.5
18
3973.1 3939.2 2899.6 2945.5 4053.7
19
3948.4 3983.1 2910.6 2958.4 4037.6
20
3964.2 3941.6 2909.1 2959.3 4039.8

P6
(kPa)
2996.5
2967.9
2953.5
3006.3
3085.4
3019.1
3269.3
3200.0
3166.1
3245.3
3297.1
3295.4
3400.8
3347.9
2960.3
2918.3
2951.1
2913.4
2900.5
2912.2

115
Table E-4: Experimental flow rate data.
Run
mr
mw
mg
index (g/s) (g/s)
(g/s)
1
85.95 69.14 154.06
2
85.15 67.57 153.97
3
83.63 64.17 153.76
4
84.44 63.86 154.07
5
86.09 66.13 154.26
6
85.09 68.53 154.27
7
90.27 70.19 154.04
8
91.43 69.39 154.38
9
90.64 70.35 154.47
10
62.31 47.90 156.56
11
62.50 48.06 156.68
12
62.45 48.19 156.65
13
62.17 45.94 156.49
14
62.60 45.59 156.42
15
82.51 58.54 149.60
16
82.64 58.42 149.95
17
83.10 58.33 149.74
18
83.69 58.91 149.82
19
83.56 58.91 150.22
20
83.35 58.50 149.72

116
Table E-5: Experimental power, performance and heat transfer data.
Run Index
r
w
QD
QA W pump Wturb II
(-)
(-)
(kW) (kW) (kW) (kW) (%)
1
0.445 0.323 5.94 5.84 0.084 0.098 1.19
2
0.445 0.312 5.93 5.82 0.094 0.108 1.18
3
0.445 0.284 5.93 5.80 0.099 0.129 2.56
4
0.445 0.281 5.97 5.84 0.092 0.123 2.12
5
0.445 0.296 5.92 5.82 0.103 0.105 0.18
6
0.445 0.321 5.93 5.82 0.086 0.102 1.34
7
0.445 0.299 7.14 6.99 0.135 0.145 0.66
8
0.445 0.291 7.10 6.94 0.130 0.159 1.91
9
0.445 0.286 7.16 6.99 0.130 0.170 2.55
10
0.445 0.304 5.07 5.02 0.044 0.051 0.65
11
0.445 0.291 5.08 5.03 0.043 0.051 0.74
12
0.445 0.294 5.07 5.02 0.043 0.052 0.90
13
0.445 0.278 5.09 5.04 0.045 0.052 0.69
14
0.445 0.268 5.09 5.04 0.040 0.052 1.16
15
0.545 0.382 5.13 4.82 0.166 0.310 8.81
16
0.545 0.381 5.13 4.83 0.149 0.302 9.73
17
0.545 0.372 5.13 4.81 0.180 0.320 8.45
18
0.545 0.375 5.13 4.80 0.175 0.326 8.52
19
0.545 0.375 5.13 4.81 0.180 0.318 8.67
20
0.545 0.372 5.13 4.81 0.171 0.317 8.51

Potrebbero piacerti anche