Download as pdf or txt
Download as pdf or txt
You are on page 1of 97

UNIVERSITY OF CALGARY

Preliminary Application Analysis of Supercritical CO2 Technology for Power Generation

in The Energy Industry of Alberta

by

Junan Rao

A RESEARCH PROJECT SUBMITTED

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN SUSTAINABLE ENERGY DEVELOPMENT

CALGARY, ALBERTA

August, 2018

© Junan Rao 2018


Abstract

sCO2 power cycle is an emerging technology for power generation and has the potential
to help meet Alberta’s future energy demand and significantly reduce GHG emissions in
power generation. To meet the Alberta Government’s 2030 energy projection, the
government should not only focus on the development of natural gas-fired power facilities,
but also consider adopting sCO2 technology. My research compares a NGCC w/CC
power plant using post-combustion technology with a direct natural gas-fired oxyfuel
sCO2 power plant in terms of energy, environmental and economic performance as the
three analytical pillars. Alberta’s climatic and industrial conditions are taken into
consideration for adopting the technology in Alberta. Potential applications of sCO2
technology in several energy sectors is also explored. The results show this emerging
technology is better and both technologically and economically feasible to be adopted in
Alberta as an innovative pathway to achieving the Alberta’s goal of a low carbon future.

ii
Acknowledgements

My most sincere appreciation and gratitude goes to my supervisor, Dr. Maryam MKhani.
This research project would not have been possible without all the supports and helps
she provided throughout this journey. Also, with her personal network within the energy
industry, I was fortunate to connect with Canadian Natural Resource Limited (CNRL) and
Gas Technology Institute (GTI) as my two industry advisors to provide industrial insights
on the sCO2 technology and its current R&D progress. The time she gave to me, and the
willingness to share her thoughts, will not be forgotten.

My sincere appreciation and gratitude goes to Dr. Mohamed Elgarni, process Engineer
for, one of the largest independent crude oil and natural gas producers in Upgrading &
Utilities Division for the Horizon Oil Sands, and Dr. Scott Macadam, Institute Engineer for
the Gas Technology Institute, an organization that leads research, development and
training in addressing energy and environmental challenges in the United States. The
knowledge shared was invaluable and significantly helpful in constructing my research.
Their willingness to share industrial R&D information on the technology was much
appreciated. I would also like to thank Ron Leoffler and Dennis Elias from Enmax for
providing insightful information on NGCC of SEC that are under public domain.

This research project would not have been possible without the support of Dr. Irene
Herremands. I thank her for all the time, enthusiasm and passion she puts into this course,
and the dedication she shows towards the students. Also, I would like to take this
opportunity to thank my family for all the support, love and encouragement they have
given me throughout this journey.

iii
Table of Content

Approval Page .................................................................................................................. i

Abstract ............................................................................................................................ii

Acknowledgements ......................................................................................................... iii

Table of Content ..............................................................................................................iv

List of Tables ................................................................................................................. viii

List of Figures ..................................................................................................................ix

List of Abbreviations and Acronyms ................................................................................xi

Chapter 1. Introduction .................................................................................................... 1

1.1 Electricity Market Transition ................................................................................... 2

1.2 Alberta Renewable Energy Potentials ................................................................... 2

1.3 Future Energy Profile of Alberta ............................................................................ 3

1.4 Emerging Power Generation System..................................................................... 4

1.5 Carbon Capture Technology in Alberta.................................................................. 5

1.6 Research Intention................................................................................................. 6

Chapter 2. Technology Review ....................................................................................... 8

2.1 NGCC for Power Generation ................................................................................. 8

2.2 Post-combustion Technology .............................................................................. 10

2.3 Oxyfuel Combustion Technology ......................................................................... 12

2.4 sCO2 Brayton Cycle for Power Generation .......................................................... 13

2.4.1 Characteristics and Advantages .................................................................... 14

2.4.2 Research and Developments ........................................................................ 17

2.4.2.1 Recompression Cycle ............................................................................. 17

2.4.2.2 Indirect-heated sCO2 Recompression Power Cycle ............................... 18

iv
2.4.2.3 Direct-heated Oxyfuel sCO2 Power Cycle .............................................. 21

Chapter 3. Methodology & Objectives ........................................................................... 24

3.1 Technology Definitions ........................................................................................ 25

3.1.1 NGCC Systems ............................................................................................. 25

3.1.1.1 Power Generation System 1 ................................................................... 26

3.1.1.2 Power Generation System 2 ................................................................... 27

3.1.2 Direct Natural Gas-fired Oxyfuel sCO2 Power Plant ...................................... 29

3.1.2.1 Power Generation System 3 ................................................................... 30

3.2 Comparison Scenario .......................................................................................... 32

3.2.1 General Assumptions and Parameters.......................................................... 33

3.3 Objectives ............................................................................................................ 34

Chapter 4. Comparison – Energy Performance ............................................................ 35

4.1 Assumptions and Parameters.............................................................................. 35

4.1.1 PGS2............................................................................................................. 35

4.1.2 PGS3............................................................................................................. 35

4.2 Evaluation & Discussion ...................................................................................... 37

Chapter 5. Comparison – Environmental Performance ................................................. 40

5.1 Assumptions and Parameters .............................................................................. 40

5.1.1 Raw Water Usage ......................................................................................... 40

5.1.2 CO2 Emission ................................................................................................ 40

5.2 Evaluation & Discussion ...................................................................................... 41

5.2.1 Raw Water Usage ......................................................................................... 41

5.2.2 CO2 Emission ................................................................................................ 42

Chapter 6. Comparison – Economic Performance ........................................................ 44

6.1 Costs Estimating Methodology ............................................................................ 44

v
6.1.1 Assumptions and Parameters ....................................................................... 44

6.1.1.1 Capital Cost ............................................................................................ 44

6.1.1.2 sCO2 Plant Components Cost Estimates ................................................ 46

6.1.1.3 O&M Costs ............................................................................................. 47

6.1.1.4 COE Estimate ......................................................................................... 48

6.2 Evaluation & Discussion ...................................................................................... 48

Chapter 7. Alberta Considerations ................................................................................ 52

7.1 Energy & Environmental Performances ............................................................... 52

7.1.1 Climatic Condition ......................................................................................... 52

7.1.1.1 Ambient Condition .................................................................................. 52

7.1.2 Industrial Condition........................................................................................ 54

7.1.2.1 Oxygen Supply in Alberta ....................................................................... 54

7.2 Economic Performance ....................................................................................... 55

7.2.1 Climatic Condition ......................................................................................... 55

7.2.2 Industrial Condition........................................................................................ 56

7.2.2.1 Regional Adjustments ............................................................................. 56

7.2.2.2 Generation Location ............................................................................... 56

7.2.2.3 Carbon Tax ............................................................................................. 57

Chapter 8. Potential Applications in Alberta .................................................................. 58

8.1 Coal Generation Sector ....................................................................................... 58

8.2 Solar Energy Sector............................................................................................. 59

8.3 Biomass to Energy............................................................................................... 61

8.4 Waste Heat to Energy.......................................................................................... 61

Chapter 9. Conclusions ................................................................................................. 63

9.1 Results of Research ............................................................................................ 63

vi
9.2 Summary ............................................................................................................. 64

References .................................................................................................................... 66

Appendix A – Main Sections and Respective Major Components of PGS2 and PGS3. 76

Appendix B – Assumed Environmental Conditions ....................................................... 77

Appendix C – Main Assumptions & Parameters of PGS2 for the Energy Comparison . 78

Appendix D – Main Assumptions & Parameters of PGS3 for the Energy Comparison . 79

Appendix E – Main Assumptions & Parameters for the Economic Comparison ............ 80

Appendix F – CO2 Emission Calculation....................................................................... 81

Appendix G – Detailed of Total Plant Cost Estimates for PGS2 .................................... 82

Appendix H – Selected Cost Estimates Information for sCO2 Power Plant ................... 84

Appendix I – COE Summary Calculation of Both Power Plants .................................... 85

vii
List of Tables

Table 1. The specific heat capacity of CO2 at different temperature and pressure. ...... 15

Table 2. Overview of NGCC w/o CC cases from studies of EBTF (2011) & NETL (2015)
and the NGCC power plant of SEC w/o CC. ................................................................. 25

Table 3. Natural gas composition from projects of EBTF (2011) & NETL (2015) against
Alberta’s natural gas composition reported by AER in 2018. ........................................ 26

Table 4. Power summary of the two power plants. ........................................................ 37

Table 5. Energy performance summary for the two simulated power plants. ................ 38

Table 6. Parameters of water balance for both power plants. ....................................... 41

Table 7. Raw water usage summary for both power plants........................................... 42

Table 8. Parameters of carbon balance for both power plants. ..................................... 42

Table 9. CO2 emission summary for both power plants. ............................................... 43

Table 10. List of final cost estimates of outstanding components differential between the
two power plants. .......................................................................................................... 49

Table 11. Final cost estimates and COE summary of selected aspects for both power
plants............................................................................................................................. 50

Table 12. Uncertainty applied to the total cost estimates and COE for both power plants.
...................................................................................................................................... 51

viii
List of Figures

Figure 1. Current and AESO’s reference case of projected future energy profile in Alberta.
........................................................................................................................................ 3

Figure 2. The four carbon dioxide capture pathways....................................................... 5

Figure 3. The simplified operational process of a typical conventional natural gas-fired


combined cycle for power generation. ............................................................................. 9

Figure 4. Flow diagram of the operational process of an aqueous amine CC technology.


...................................................................................................................................... 11

Figure 5. A typical process of oxyfuel combustion. ....................................................... 13

Figure 6. The operational process of a state-of-the-art natural gas-fired combined cycle


for power generation. .................................................................................................... 14

Figure 7. The size of a sCO2 turbine compared with a conventional steam turbine under
an equal rated power output. ......................................................................................... 16

Figure 8. The size of a direct oxyfuel sCO2 plant against a natural gas-fired power plant
and a coal-fired power plant. ......................................................................................... 16

Figure 9. The general diagram of an indirect-heated sCO2 recompression power cycle.


...................................................................................................................................... 18

Figure 10. Diagram of the developing 10 MW test facility with indirect solids heated sCO2
recompression power cycle. .......................................................................................... 19

Figure 11. The general diagram of a basic direct-heated sCO2 power cycle. ................ 21

Figure 12. Basic diagram of the developing 50 MW test facility with direct natural gas
heated sCO2 power cycle. ............................................................................................. 22

Figure 13. Demonstration of flow process of the NGCC power plant of SEC w/o CC. .. 27

Figure 14. Demonstration of flow process of the NGCC power plant of SEC w/CC. ..... 28

ix
Figure 15. Flow diagram of the direct natural gas-fired oxyfuel sCO2 power plant. ....... 31

Figure 16. The designed model of sCO2 turbine by General Electric Global Research. 36

Figure 17. Each capital cost level and its respective elements. .................................... 44

Figure 18. A 25 MW sCO2 oxy-turbine of Toshiba......................................................... 46

x
List of Abbreviations and Acronyms

AESO Alberta Electric System Operator


ANPO Annual Net Power Output
ASU Air Separation Unit
BEC Bare Erected Cost
CC Carbon Capture
CCF Capital Charge Factor
CCS Carbon Capture & Storage
CF Capacity Factor
CNRL Canadian Natural Resource Limited
CO Carbon Monoxide
COSIA Canada’s Oil Sand Innovation Alliance
CPU Carbon Processing Unit
CSP Concentrated Solar Plant
EBTF European Benchmarking Task Force
EPC Engineering, Procurement and Construction
EPCC Engineering, Procurement and Construction Cost
EOS Equation of State
FC Fuel Cost
GHG Greenhouse Gas
GTI Gas technology Institute
HP High Pressure
HRSGs Heat Recovery Steam Generators
HTR High Temperature Recuperator
IP Intermediate Pressure
ITR Intermediate Temperature Recuperator
KPI Key Performance Indicator
xi
kW kilowatt
kWh kilowatt-hour
COE Cost of Energy
LP Low Pressure
LTR Low Temperature Recuperator
MC Main Compressor
MHE Main Heat Exchanger
MR Main Recuperator
MW megawatts
MWh megawatt-hour (annual)
NETL National Energy Technology Laboratory
NGCC Natural Gas-fired Combined Cycle
O&M Operation & Maintenance
PGS Power Generation System
RC Recompressing Compressor
sCO2 Supercritical Carbon Dioxide
SEC Shepard Energy Centre
TASC Total As-Spent Cost
TFC Total Fix O&M Cost
TIT Turbine Inlet Temperature
TOC Total Overnight Cost
TOT Turbine Outlet Temperature
TPC Total Plant Cost
TVC Total Variable O&M Cost
w/CC with Carbon Capture
w/o CC without Carbon Capture
WHR Waste Heat Recovery

xii
Chapter 1. Introduction

Under the Paris Agreement, Canada is committed to reducing its greenhouse gas (GHG)
emissions, and the federal government has requested individual provinces to develop
their own strategy and meet the national GHG emission reduction target collectively by
2030 – the target being 30% below the 2005 level (Government of Canada, 2018). In
2015, Alberta announced its Climate Leadership Plan which includes implementing
carbon taxes, cutting methane emission, more renewable energy development, capping
oil sand emissions, and phasing out coal power plants (Alberta Government, n.d.a). In
2016, the provincial government announced its plan to shift Alberta’s energy-only market
system to capacity and energy market, and by 2030, 70% of electricity will come from
natural gas and 30% will be generated by renewable energy (Alberta Government, 2017;
Alberta Electric System Operator, 2017).

Currently, Alberta’s energy makeup is dominated mostly by coal with natural gas following
closely behind. Other energy sources include hydro, wind, solar and biomass as
renewables. As coal power plants are phasing out gradually, the provincial government
plans to build more low-emission power plants along with renewables to offset the loss of
the province’s energy capacity. Currently, the most energy efficient power generation
system is a natural gas-fired combined cycle (NGCC) with up to 60% plant efficiency
(Jarre, Noussan, & Poggio, 2016). In the meantime, carbon capture (CC) was identified
as one of the key CO2 mitigation tools to meet the province’s Climate Change Strategy
Goal. To demonstrate environmental leadership and decrease GHG emissions, the
Government of Alberta introduced policies that would incentivize the energy industry for
CC technology development, and along with the NRG COSIA Carbon XPRIZE
competition hosted by Canada’s Oil Sand Innovation Alliance (COSIA), encouraged the
global development of carbon conversion technology.

Thus, to meet the 70% requirement, the Alberta Government is planning to build more
natural gas-fired power plants that may possibly be equipped with CC technology in the
future to meet the growing energy demand and emission reduction target (Alberta Electric
System Operator, 2017). However, considering the intermittency issue of renewables,

1
many argue that meeting the energy demand is questionable. Fortunately, in recent years,
a new power generation system — supercritical carbon dioxide (sCO2) Brayton power
cycle — emerged in the energy industry after years of research and it is believed to be
the next era of power generation that has near zero carbon emission.

1.1 Electricity Market Transition

Energy-only market and capacity market are the two main electricity market systems
existing globally. Alberta’s current energy-only electricity market system has been facing
multiple issues for the past several years. Consumers are vulnerable to volatilities in
electricity pricing, which is also less attractive to green energy investors in Alberta’s
electricity market, due to the market failing to provide a stable and reliable electricity
source (Olmstead & Ayres, 2014; Alberta Government, 2016). Meanwhile, Alberta is
falling behind the pace of a global movement that is moving toward low-carbon and
sustainable electricity (Alberta Government, 2016). To overcome these issues and ensure
a reliable electricity supply to consumers, the Alberta government announced to shift the
current electricity market system to capacity and energy market (Alberta Electric System
Operator, 2017). With the implementation of the Alberta Climate Leadership Plan carried
out in 2016, coal power plants are determined to be phased out completely in the near
future, and the provincial government is projecting that 70% of electricity will come from
natural gas and 30% will come from renewable energy by 2030 (Alberta Electric System
Operator, 2017).

1.2 Alberta Renewable Energy Potentials

Onshore wind energy is currently the major renewable source for Alberta. Geographically,
all the highest onshore wind potentials are located in the southeastern region of Alberta,
despite environmental constraints, and theoretically, it is estimated that 25% of the
potentials are subjected to physical availability and economic feasibility (Barrington-Leigh
& Ouliaris, 2017) with a total of 3200 MW potential installed capacity (Energy Resolution
Services, 2016). Solar energy is still a relatively new renewable energy source in the
Alberta electricity market, and due to the occasional market power exercise and price
volatility, could be the reason why some solar energy investors are unwilling to participate
2
in Alberta’s current electricity market (Olmstead & Ayres, 2014). Like wind energy, the
southeast areas of Alberta have the highest solar farm potentials with a total potential
installed capacity of 5000 MW (Energy Resolution Services, 2016; Barrington-Leigh &
Ouliaris, 2017). The government of Alberta desires to develop more of these two energy
sources to offset the loss of 67% of the existing coal generation capacity (Layzell,
Shewchuk, Sit, & Klein, 2016).

1.3 Future Energy Profile of Alberta

According to the Alberta Electric System Operator (AESO), the total installed natural gas
capacity in Alberta in 2017 was 7596 MW, and 10% of that is combined cycle. Other
natural gas-fired technologies, such as cogeneration and single cycle are 30% and 5%,
respectively.

Figure 1. Current and AESO’s reference case of projected future energy profile in Alberta.

(Alberta Electric System Operator, 2017)

Figure 1 above shows the current and projected future energy profile of Alberta. Although
the total natural gas capacity for power generation is expected to increase to nearly
14,900 MW, the percentage of combined cycle and single cycle in the province are
projected to be increased to 27% and 11 %, respectively, in contrast to a reduction to 22%
for cogeneration in the next 20 years. The NGCC will be the main low-emission, most

3
efficient, and fossil fuel-based power generation development in the future for Alberta to
meet the 70% projection (Alberta Electric System Operator, 2017; Layzell et. al., 2016).
For example, a 1060 MW state-of-the-art NGCC power plant, a joint venture project
between Capital Energy and Enmax, will commence in 2019 (Alberta Government, n.d.b).
Also, cogeneration and single cycle will be developed together to combat the
intermittency challenge of renewable energy.

1.4 Emerging Power Generation System

sCO2 Brayton power cycle for power generation has been studied by researchers for the
past decade. It has been the focus of increased attention in the past three years in
particular for the global energy industry. The technology is proved to be transformative
with the highest potential to be the next era of power generation. sCO2 power cycle has
a wide range of applications, such as waste heat recovery (WHR), concentrated solar
power (CSP), nuclear power, and fossil fuel combustion (Ahn, et al., 2015; Energy, 2015).
Compared with any current existing Rankine and Brayton cycle technologies, sCO2 power
cycle possesses advantages over other systems, such as higher electric efficiency, lower
capital and operating costs, lower environmental footprint, and near zero GHG emissions
(Energy, 2015; Ahn, et al., 2015). Recently, a number of developments of sCO2 power
cycle in pilot scale are emerging. The world’s first pilot project of indirect natural gas-
heated sCO2 recompression power cycle test facility is currently under development,
which is sponsored by the US. Department of Energy, in San Antonio, Texas (US.
Department of Energy, n.d.). In the meantime, a joint-venture project of direct natural gas-
fired sCO2 power cycle, or Allam Cycle test facility, is currently in continuous operation in
La Porte, Texas (Allam, et al., 2017). There are also a variety of developments for different
applications in Europe, Asia, and Australia (Crespi, Gavagnin, Sanchez, & Martinez,
2017).

To my knowledge, no relevant research and development has been conducted to the


same extent in Canada. Natural Resource Canada in conjunction with Carleton University,
however, has developed a transformative, patented, and fossil-fueled combustion
technology, known as G2 technology, and it is a highly pressurized and oxyfuel

4
compacted combustion system (National Energy Technology Laboratory, 2016). The G2
technology supports sCO2 power cycle for high cycle efficiency, high purity of CO 2, and
sCO2 turbine size reduction (National Energy Technology Laboratory, 2016). Oxy-
combustion is a necessity for sCO2 power cycle. The technology is not considered to be
a carbon capture technology but rather as an inherent function because the sCO2 power
cycle uses CO2 as the working fluid to generate power.

1.5 Carbon Capture Technology in Alberta

Carbon capture technology can be divided into four different types of systems in relation
to how it captures CO2. They are post-combustion, pre-combustion, oxyfuel combustion,
and industrial process separation (Natural Resource Canada, 2016a).

Figure 2. The four carbon dioxide capture pathways.

(Natural Resource Canada, 2016a)

5
The above figure shows the simplified process for each technology. Captured CO2 can
be utilized in many ways, such as enhanced oil recovery, enhanced coal bed methane
recovery, enhanced geothermal systems, and stored permanently in underground
storage (Global CCS Institute, n.d.). The feasibility of CC technology has already been
demonstrated in Alberta. Shell’s Quest project is Alberta’s first full-scale carbon capture
and storage project and operates at an oil sands upgrader facility in Fort Saskatchewan,
Alberta. The project adopts post-combustion technology using Cansolv-Amine as its CC
technology. Also, a number of oil & gas organizations have adopted the technology for
certain industrial activities. For instance, CNRL incorporates pre-combustion technology
to separate and recover CO2 from syn-gas in their hydrogen plant at their oil sand mining
facilities and inject the captured CO2 into their tailings ponds to further consolidate fluid
tailings (Dr. M. Elgarni, personal communication, February 28, 2018).

As Figure 2 demonstrates, pre-combustion capture technology focuses on capturing the


CO2 prior to combustion completion. This is done primarily through the gasification
process where the fuel source is converted to hydrogen (H2) and carbon monoxide (CO),
which is known as syngas. The syngas then goes through a steam reformer, undergoing
a water-gas shift reaction where CO and H2O are converted into CO2 and H2. Hydrogen
is then combusted as the fuel source which yields only H2O in the flue gas. The CO2
derived from the steam reformer is then dehydrated and compressed for transportation
and storage (Leung, Caramanna, & Maroto-Valer, 2014).

1.6 Research Intention

To meet the provincial government’s 2030 energy projection for emission reduction target,
the Alberta government and the energy industry should not only invest in renewable
energy, mainly wind and solar (Alberta Government, n.d.c; Energy Resolution Services,
2016), and natural gas-fired power plants with CC technology, particularly NGCC
developments (Alberta Electric System Operator, 2017), but also in sCO2 technology. As
a world leader in energy, the Alberta government should consider adopting this promising
technology as an innovative pathway to ensuring that future energy demand is met both
economically and environmentally.

6
I am conducting this research to explore and evaluate the sCO2 power cycle technology
in terms of Energy, Environmental and Economic performances, as the three research
pillars for my capstone research project, and, in comparison with NGCC technology to
get a better understanding of its performance and application in the energy industry,
specifically for power generation in Alberta. Chapter 3 – Methodology and Objectives,
illustrates details of the specific aspect which will be analyzed for each pillar and how the
comparisons are carried out.

7
Chapter 2. Technology Review

Technologies that will be considered in my research are reviewed through literature and
personal communications with industry representatives. As per Chapter 1, reviews of
NGCC, CC technologies, and sCO2 power cycles are carried out in this Chapter.
Meanwhile, post-combustion can be retrofitted into existing plants with significantly fewer
modifications and costs than pre-combustion and is more favored in power generation
(Natural Resource Canada, 2018; Subramanian, Jordal, Anantharaman, Hagen, &
Roussanaly, 2017). Study of White & Weiland (2018) has used post-combustion
technology for a state-of-the-art NGCC power plant to compare with a direct natural gas-
fired oxyfuel sCO2 power plant. In addition, oxyfuel combustion unit is a necessity for a
direct sCO2 power cycle to achieve optimization and produce high purity and
concentration of CO2. Therefore, only CC technology of post-combustion and oxyfuel
combustion are considered in my research.

2.1 NGCC for Power Generation

Combined cycle gas turbine and single cycle gas turbine are the two types of natural gas-
fired generation systems. Combined cycle can serve as based-load generation or for
peak-load demand thanks to its technological maturity and reliable power supply, and it
is more thermally efficient than single cycle. Combined cycle power generation is a typical
combination of Brayton and Rankine thermodynamic cycles to generate power. NGCC
has both gas turbine and steam turbine for electric generation. Gas turbine serves as the
topping cycle and steam turbine is the bottoming cycle in NGCC. Individually, steam
Rankine cycle has a higher efficiency than simple cycle gas turbine, but it requires
intensive water consumption for power generation. By combining these two cycles
together, the overall efficiency for power generation can be greatly increased. The
efficiency of an average NGCC ranges from 45 – 52.5% (Penkuhn & Tsatsaronis, 2016),
however, a state-of-the-art NGCC has the highest net thermal efficiency that could reach
up to 60% (Jarre, Noussan, & Poggio, 2016) among all the existing power generation
technologies available.

8
Figure 3. The simplified operational process of a typical conventional natural gas-fired
combined cycle for power generation.

(Hislop, 2016)

The above is a simplified description of an NGCC and Figure 3 is the associated


operational process diagram. For a typical conventional NGCC power plant, air is
compressed into combustor while injecting natural gas for combustion. The hot exhaust
gas then drives the gas turbine(s) for generating power through expansion. After the first
electric generation through fuel combustion, the heat from high temperature exhaust
gases coming out of the gas turbine is captured by employing a series of heat recovery
steam generators (HRSGs) or a boiler to heat up water and generate steam in tubes.
Normally, natural circulation or pumping provides a flowing stream among HRSGs for
steam generation. The HRSGs includes various modules that serve different purposes in
the process of dry superheated steam production. The water is preheated to the boiling
point in the heat exchanger (economizer) and flows in tubes (evaporator) where hot
exhaust gases release heat and steam is generated. At last, superheater further heats
9
the saturated steam to dry steam. The steam is then supplied to the steam turbine where
it expands and imparts turbine rotation to produce electricity.

The NGCC power plant of Shepherd Energy Centre (SEC), located south-east of Calgary,
is a state-of-the-art NGCC power plant which is owned and operated by Enmax
Corporation. Its capital cost is around C$1.4bn (R. Loeffler, personal communication,
January 19, 2018). The plant has been operational since 2015 and served as a base load
generator. The configuration of the power plant is two gas turbines (240 MW each) by
one steam turbine (360 MW) with two Vogt triple pressure HRSGs installed (Mitsubishi
Hitachi Power Systems Global, n.d.). The power plant has a capability of reaching a
turbine inlet temperature (TIT) of 1600°C and 600°C as the turbine outlet temperature
(TOT) (R. Loeffler, personal communication, January 19, 2018; Mitsubishi Hitachi Power
Systems Global, n.d.). It is the largest NGCC power generation facility in Alberta with a
capacity of 800 MW and can serve more than half of the current energy demand of
Calgary (Power Technology, n.d.). The flue gas is only treated with ammonia for air
emission reduction, and the plant generates low concentrations of SO x and NOx (R.
Loeffler, personal communication, January 19, 2018). Furthermore, the facility has the
ability to reduce carbon emission by about half compared to that of a conventional coal-
fired power plant (Power Technology, n.d.). It is the first power plant in Alberta that uses
reclaimed water for power generation with a requirement of approx. 14 million litres per
day (Enmax, 2015). Interestingly, the SEC was selected for the COSIA’s carbon XPRIZE
global carbon conversion technology competition in 2015 (Enmax, n.d.). The power plant
has a net plant efficiency of about 60% (R. Loeffler, personal communication, January 19,
2018; Mitsubishi Hitachi Power Systems Global, n.d.). Therefore, the SEC has a state-of-
the-art NGCC that doesn't have the function of capturing CO2. The concentration of CO2
in the flue gas is about 4% (National Energy Technology Laboratory, 2015).

2.2 Post-combustion Technology

Post-combustion is considered to be the prevailing technology used in retrofitting existing


power plants as well as in newly constructed plants (Leung, Caramanna, & Maroto-Valer,
2014) due to its economic and technological feasibility as an “end-of-pipe” solution (Wang,

10
Zhaoa, Ottoa, & Robinius, 2017). It is a process of removing CO2 from flue gas after the
combustion of fossil fuel. The major concerns for deploying post-combustion technology
in large-scale power plants include waste disposal (e.g. degraded amine waste), the type
of solvent employed, and overall efficiency of the process (Leung, Caramanna, & Maroto-
Valer, 2014). The cost of carbon transportation and storage, as well as energy penalty or
parasitic load associated with CO2 capturing and compression, are significantly
dependent on the overall efficiency of the technology (Leung, Caramanna, & Maroto-
Valer, 2014) which could result in higher capital cost and cost of energy (COE) (Penkuhn
& Tsatsaronis, 2016).

Figure 4. Flow diagram of the operational process of an aqueous amine CC technology.

(National Energy Technology Laboratory, 2014)

Four common post-combustion industrial technologies are identified as aqueous amine,


polymeric membranes, a low temperature sorbent and Ca-loopping (Subramanian et al.,
2017; Natural Resource Canada, 2018). Among these four technologies, aqueous amines
is deemed to be the best option for CC for NGCC thanks to the technology maturity and
relatively simple process integration (Subramanian, et al., 2017), and it has been
considered as CC technology for NGCC in research studies such as National Energy
Technology Laboratory (NETL) (2015) and European Benchmarking Task Force (EBTF)
(2011). The type of amine used in this technology is Methyl-Ethanolamine or MEA
11
(European Benchmarking Task Force, 2011; Subramanian, et al., 2017; White & Weiland,
2018; National Energy Technology Laboratory, 2015). The engineering process of the
technology is shown in Figure 4. After the exhaust gas exits the HRSGs and is cooled, it
flows into an absorption tower that fills with the liquid amine and CO2 from the exhaust
gas is absorbed by the countercurrent liquid amine (Technology Centre Mongstad, 2010).
The CO2 saturated amine solvent is then pumped to a regenerator to release CO2 through
chemical reaction under a condensing steam flowing upwards (Technology Centre
Mongstad, 2010; Subramanian, et al., 2017). The released CO 2 is then compressed and
ready for transport and storage.

2.3 Oxyfuel Combustion Technology

The development and commercial demonstration of the oxyfuel combustion technology


has been recognized as an emerging opportunity for designing and developing a new
power cycle that has higher efficiency, lower environmental footprints, and full scaled CO 2
capture (Natural Resource Canada, 2016b), such as sCO2 power cycle. For example, the
G2 technology is a highly pressurized and oxyfuel compacted combustion system
(National Energy Technology Laboratory, 2016).

In the oxyfuel combustion process, the nitrogen in ambient air is eliminated to yield pure
oxygen as the oxidant in fossil fuel combustion. This is achieved through the integration
of an Air Separation Unit (ASU). Cryogenic method is the predominant technology in
producing high oxygen thanks to its ability for producing greater than 99% oxygen purity
and significantly higher capacity, but requiring considerably higher energy consumption
(Chorowski & Gizicki, 2015). The method utilizes condensation temperature differential in
specific gases to accomplish the goal of separation (Chorowski & Gizicki, 2015).

In case of a coal-fired power plant, after particulates and SO2 are removed through
electrostatic precipitation and desulphurization methods, CO2 and water become the
primary components of flue gas (Leung, Caramanna, & Maroto-Valer, 2014). The flue gas,
containing a high concentration of CO2, can be compressed through a direct compression
method or by applying a low-temperature separation method (Natural Resource Canada,

12
2016a). Figure 5 illustrates a more detailed process of oxyfuel combustion than that of
Figure 2.

Figure 5. A typical process of oxyfuel combustion.

(Natural Resource Canada, 2016b)

Oxyfuel combustion provides significantly high-purity oxygen content for fuel combustion
which results in a high concentration of CO2 and a possible size reduction of the
combustor compared with an air-fired combustor (Natural Resource Canada, 2016a). A
drawback for this technology is the potential for corrosion due to the possibility of high
SO2 concentration in flue gas within the system (Leung, et al., 2014). Additionally, it
requires an intensive gas separation unit to support high volumetric pure oxygen
consumption for the process (Leung, et al., 2014) which can significantly decrease the
overall efficiency of the respective power plant and increase the plant’s capital cost and
COE, and this effect is higher compared with that of post-combustion (Penkuhn &
Tsatsaronis, 2016).

2.4 sCO2 Brayton Cycle for Power Generation

Brayton cycle is an ideal cycle for a gas-turbine engine, and it implements both the
processes of compression and expansion. sCO2 power cycles employ supercritical CO2
as the medium in a thermodynamic process of Brayton cycle. In the past decades, sCO2
13
power cycles have been investigated, and recently, have received considerable attention
in the global energy industry due to its wide range of applications, from high temperature
in fossil fuel combustion to low temperature in geothermal (Ahn, et al., 2015; Zhu, 2018;
Energy, 2015).

2.4.1 Characteristics and Advantages

Figure 6. The operational process of a state-of-the-art natural gas-fired combined cycle


for power generation.

(Energy, 2015)

Figure 6 shows the physical properties of CO2 under different pressures and temperatures.
Carbon dioxide, as a working fluid, has a low critical pressure and temperature points of
74 bar and 31oC (Ahn, et al., 2015; Zhu, 2018; Energy, 2015), and can be easily
compressed and heated to the critical state under adequate conditions. Supercritical CO2
is carbon dioxide at its critical state where liquid and gas densities are the same (US.
Department of Energy, 2015; Park, Kim, Yoon, Rhim, & Yeom, 2018; Ahn, et al., 2015).
14
sCO2 has several characteristics. Gas is more compressible than liquid, and it requires
higher energy input during the compression process (Zhu, 2018; Ahn, et al., 2015). This
explains why the thermal efficiency of a gas turbine is lower than a steam Rankine cycle
(Ahn, et al., 2015). CO2, as a liquid, becomes less compressible near its critical point
which significantly reduces the energy consumption of compression, and it results in high
cycle efficiency (Zhu, 2018; Energy, 2015). The efficiency increases with rising inlet
turbine temperature (Energy, 2015). In sCO2 power cycle, the fluid remains in high density
and has a low pressure ratio due to its decreased critical temperature and pressure points
(Ahn, et al., 2015). Unlike conventional gas turbine working fluid, the correlation between
pressure and the specific heat capacity of CO2 is positive (Penkuhn & Tsatsaronis, 2016;
Scaccabarozzi, Gatti, & Martelli, 2016; White & Weiland, 2018). To summarize, high
temperature and low pressure CO2 absorbs less thermal energy due to lower specific
heat capacity in contrast to the higher specific heat capacity of low temperature and high
pressure CO2. As Table 1 illustrates below, the specific heat capacity of CO2 decreases
under constant pressure between the temperature of 350 to 750K (Scaccabarozzi et al.,
2016).

Table 1. The specific heat capacity of CO2 at different temperature and pressure.

CO2 specific heat capacity under constant pressure (kJ/kg.K)


Temperature (K) 100 Bar 200 Bar 300 Bar
350 1.96 2.57 2
400 1.32 1.86 1.9
450 1.2 1.47 1.63
500 1.17 1.34 1.47
600 1.16 1.24 1.31
750 1.18 1.21 1.23
1000 1.25 1.27 1.28

(Scaccabarozzi et al., 2016)

High power density is achieved due to the working fluid which has a higher density (twice
denser than steam) and volumetric heat capacity (similar viscosity as gas) in the
supercritical state. This results in a significant reduction of component sizes within the
power cycle system, such as turbine (about 10 times smaller than conventional power
15
turbine, Figure 7) and heat exchangers (Zhu, 2018; Ahn, et al., 2015; US. Department of
Energy, 2015; Park et al., 2018), and extraction of “pipeline-ready” CO2 (in liquid state)
without additional carbon capture equipment (Allam, et al., 2017). Consequently, the
power cycle has a high process efficiency, and it can result in a low thermal input, low
capital cost, small plant footprint (Figure 8), less water usage, and less greenhouse gas
emission (Energy, 2015).

Figure 7. The size of a sCO2 turbine compared with a conventional steam turbine under
an equal rated power output.

(Duanetildan, 2016)

Figure 8. The size of a direct oxyfuel sCO2 plant against a natural gas-fired power plant
and a coal-fired power plant.

(Net Power, n.d.)


16
2.4.2 Research and Developments

Open cycle and closed cycle are the two primary sCO2 Brayton power cycles (Energy,
2015; Zhu, 2018). Specifically, open cycle is direct-heated through a combustor, and
closed cycle is indirect-heated through a heat exchanger or exchangers. There are
several configurations of sCO2 power cycle with a various combination of recuperators
and compressors to achieve the common goal of maximum cycle and power efficiency,
and they have been reviewed by numerous researchers (Ahn, et al., 2015; Crespi, et al.,
2017). It has been proven that direct-heated cycle has a higher cycle, process, and
thermal efficiency than in-direct-heated because higher TITs can be attained when the
working fluid directly contacts with the heat source, and TITs for indirect cycles are limited
by the heat exchanger construction materials (S. Macadam, personal communication,
March 06, 2018; US. Department of Energy, 2015; Ahn, et al., 2015). The sCO2
technology is gaining popularity with many research programs and smaller scale
developments. For instance, in the US, utility-scale indirect-heated recompression and
direct-heated oxyfuel sCO2 power cycles, are developed for field testing purpose (Allam,
et al., 2017; US. Department of Energy, n.d.), which is illustrated later in this section.

On the other hand, challenges and drawbacks are foreseeable. The main challenges of
the sCO2 technology are sCO2 turbine design and sizing, material selection for the cycle,
and valve design for operation (Energy, 2015). The main drawbacks of a sCO2 power
cycle technology are technical issues regarding the handling of extreme high TIT and
plausible corrosion on the sCO2 turbine. This is due to the power cycle requiring high-
oxygen purity level (direct cycle only) and producing high concentration of CO2 (Energy,
2015; Allam, et al., 2017).

2.4.2.1 Recompression Cycle

To achieve heat loss reduction and maximize the thermal-to-power conversion, a heat
exchanger or recuperation process is necessary. In any sCO2 cycle, specific heat
capacity increases as CO2 is close to the critical point, and the thermal capacity on the
low pressure side of the recuperator where hot CO2 flows is lower than the high pressure
side of the recuperator where cold CO2 is present. Due to the fact above, the temperature
17
increase by the recuperator on the high-pressure CO2 has a limitation, and therefore this
will always result in a low cycle efficiency (Energy, 2015). Adding a recompression system
can mitigate the issue. In a configuration of an indirect-heated recompression cycle, high
and low temperature recuperators and a recompressing compressor are employed to
maximize waste heat reuse and drastically improve thermal efficiency (Ahn, et al., 2015;
Energy, 2015). Recompression cycle has been acknowledged to be the most efficient
sCO2 power cycle configuration among all possible power cycle configurations (US.
Department of Energy, 2015; Ahn, et al., 2015; Allam, et al., 2017).

2.4.2.2 Indirect-heated sCO2 Recompression Power Cycle

Figure 9. The general diagram of an indirect-heated sCO2 recompression power cycle.

(National Energy Technology Laboratory, 2017)

Figure 9 illustrates the general engineering process of the indirect-heated power cycle.
This configuration consists of a main heat exchanger (MHE), sCO2 turbine, high
temperature recuperator (HTR), low temperature recuperator (LTR), CO2 cooler, main
compressor (MC), and recompressing or recycle compressor (RC). ASU isn’t a necessity

18
for a closed cycle as the cycle indirectly contacts with a heat source or contacts at the
level of TIT.

Figure 10. Diagram of the developing 10 MW test facility with indirect solids heated
sCO2 recompression power cycle.

(National Energy Technology Laboratory, 2017)

In this configuration, the sCO2 cycles through the MHE and sCO2 turbine to generate
electricity, after which a portion of the low pressure stream existing the LTR is then
compressed by the RC to reach the maximum cycle pressure before entering the HTR.
This feature provides a better thermal capacity match between the hot and cold CO 2 and
higher effectiveness of recuperation without passing through the CO 2 cooler and low
temperature side of the LTR (Energy, 2015; Zhu, 2018; Ahn, et al., 2015). Indirect sCO2
power cycles have a wide range of potential applications including coal-fired systems,
waste heat recovery, CSP, biomass, and nuclear, but direct cycles are limited to natural
gas and syngas-fired heat sources (S. Macadam, personal communication, March 06,
2018).
19
Figure 10 shows the blueprint of the pilot project of the indirect cycle demonstration facility
in San Antonio, Texas. with an outlet turbine temperature of 700°C and an output of 10
MW (National Energy Technology Laboratory, 2017). Natural gas-fired heater was
selected as a convenient heat source for demonstrating the indirect sCO2 cycle as the
main objective of the project (S. Macadam, personal communication, March 06, 2018).
As the diagram illustrates, the working fluid, CO2, is operated to remain in the supercritical
state throughout the entire closed cycle to prevent and minimize heat loss and ensure
thermal and cycle efficiencies are maximized (National Energy Technology Laboratory,
2017). The efficiency of this pilot power plant is estimated to be 40% (S. Macadam,
personal communication, March 06, 2018) which is higher than efficiency performance of
a conventional simple gas cycle turbine of around 35% (Energy Education, 2015) and
coal-fired power plant of around 32 to 38% (World Coal Association, n.d.). Also, a goal of
the project is to demonstrate a pathway to achieving >50% cycle efficiencies for
commercial-scale indirect sCO2 plants (S. Macadam, personal communication, March 06,
2018). It is an unprecedented project, but relevant information and data are currently
unavailable in the public domain.

Indirect-heated sCO2 recompression power cycle has also been proposed and analyzed
as an alternative bottoming cycle for NGCC to replace steam Rankine bottoming cycle
(Huck, Freund, Lehar, & Peter, 2016). The study concluded that indirect-heated sCO2
bottoming power cycle does not outperform a state-of-the-art three-pressure-reheat
steam bottoming cycle of NGCC due to technical difficulty and operational safety
concerns. However, the indirect-heated sCO2 bottoming power cycle does offer an
advantage over two-pressure-non-reheat found in smaller Rankine bottoming cycles
behind aero-derivative gas turbines (Huck, Freund, Lehar, & Peter, 2016).

20
2.4.2.3 Direct-heated Oxyfuel sCO2 Power Cycle

Figure 11. The general diagram of a basic direct-heated sCO2 power cycle.

(Energy, 2015)

The general engineering process of the power cycle is demonstrated in Figure 11. In this
configuration, oxy-combustor, sCO2 turbine, main recuperator (MR), MC, and CO2 cooler
are included. Additionally, to maintain the high cycle efficiency, a consistent high purity
and concentration of sCO2 oxidant stream is required, and that can be achieved by
providing a highly purified oxygen stream via a cryogenic ASU (Energy, 2015). Normally,
unlike the indirect-heated recompression cycle, a RC, HTR and LTR are required to
maximize thermal and cycle efficiencies. However, it may change according to the power
cycle design and many direct cycles incorporate these components to achieve high plant
electric efficiency (Allam et al., 2017; Penkuhn & Tsatsaronis, 2016; Scaccabarozzi, Gatti,

21
& Martelli, 2016; Scaccabarozzi, Gatti, & Martelli, 2017; White & Weiland, 2018). As the
working fluid directly contacts with the heat source, providing significantly high TIT will
result in significantly higher cycle efficiency and higher power density (Energy, 2015).
Typically, TIT of a natural gas-fired gas turbine ranges from 1100 to 1500°C (Energy,
2015).

Figure 12. Basic diagram of the developing 50 MW test facility with direct natural gas
heated sCO2 power cycle.

(Allam, et al., 2017)

Figure 12 is the simple scheme provided by Allam, et al., (2017) of the developing 50 MW
with TIT of 1150°C test facility. This power cycle was first proposed by Allam, et al., in
2013, followed by several academic reviews, such as Penkuhn & Tsatsaronis, (2016),
Scaccabarozzi , Gatti, & Martelli, (2016), and Scaccabarozzi , Gatti, & Martelli, (2017).
These studies all have endorsed that the direct oxyfuel sCO2 power cycle of Allam has a
considerably higher efficiency than a conventional NGCC power plant. The study of Allam,
et al., (2017) claims that with the success of the 50 MW power plant by Net Power in La

22
Porte, Texas, the development of a commercial scale of 300 MW Allam cycle power plant
will be completed by 2020.

The facility uses natural gas as the heat source and requires 99.5% purity of oxygen at a
pressure of 120 bar (Penkuhn & Tsatsaronis, 2016; Scaccabarozzi et al., 2016). In this
design, after the cooling and before entering the cold side of the MR, an oxidant stream
is formed by mixing partial recycled CO2 with oxygen (Allam, et al., 2017). This stream
then enters MR to be preheated up to between 720 to 750°C (Penkuhn & Tsatsaronis,
2016; Scaccabarozzi et al., 2016), followed by the sCO2 turbine. The operating pressure
condition for oxy-combustor is between 200 to 400 bar (Scaccabarozzi et al., 2016, 2017).
The inlet turbine temperature designed for the facility is between 1100 to 1200°C
(Scaccabarozzi et al., 2016, 2017). A typical pressure designed for working fluid entering
the combustor and turbine is at least 300 bar (Allam et al., 2017; Penkuhn & Tsatsaronis,
2016; Scaccabarozzi et al., 2016, 2017; White & Weiland, 2018). Thus, gaseous fuel is
required to be compressed to around 300 bar as well (White & Weiland, 2018).

According to Allam, et al., (2017), the working fluid temperature coming out of the hot end
of the recuperator constraints the TIT. Mass balance of the process gas within the system
is a necessity, and a portion of the working fluid, which is about 5% of the total recycled
stream (Allam et al., 2017; Scaccabarozzi et al., 2016), is intentionally discharged before
MC (Allam, et al., 2017; Energy, 2015). The exported CO2 has a high purity of above 97%
(Scaccabarozzi et al., 2016) and typically leaves at a pipeline pressure of 150 bar which
reduces capital cost on additional carbon capture equipment for a traditional power plant
(Allam, et al., 2017). The CC rate is about 100% (Allam et al., 2017; Penkuhn &
Tsatsaronis, 2016; Scaccabarozzi et al., 2016). Interestingly, in this designed cycle, CO2
isn’t kept in the supercritical state entirely, but interchangeable between liquid and
supercritical states (Allam, et al., 2017). This power cycle is claimed to have 59% net
plant efficiency by Allam et al. (2017) (Scaccabarozzi et al., 2016). This is competitive to
a state-of-the-art NGCC power plant without CC technology which has an optimum net
plant efficiency of 60% (Dr. M. Elgarni, personal communication, February 28, 2018).

23
Chapter 3. Methodology & Objectives

As the review illustrates in Chapter 2, two conclusions can be drawn 1) post-combustion


with aqueous amine is a better choice for a NGCC power plant than pre-combustion and
oxyfuel combustion for capturing CO2 in terms of technological and economic feasibility;
and 2) a direct natural gas-fired oxyfuel sCO2 power plant demonstrates outperforming
potential to a NGCC power plant in terms of energy performance (net plant efficiency),
and as such a indirect-heated sCO2 recompression power plant will not be considered in
my research. The three power generation systems (PGS) identified and selected for
research are NGCC without CC technology (w/o CC) power plant of SEC (PGS1), NGCC
with CC technology (w/CC) using aqueous amine power plant (PGS2), and direct natural
gas-fired oxyfuel sCO2 power plant (PGS23). These three PGSs are further described in
this Chapter. One comparison scenario is established for my research as well as its
related necessary general assumptions and parameters for the chosen PGSs. Also, the
objectives of my research are detailed in this Chapter. The methodology of energy and
environmental performance analysis is addressed in this Chapter, and the economic
analysis methodology can be found in Chapter 6.

To perform the analysis of the chosen PGSs for comparison on the three pillars (energy,
environmental, and economic), it is necessary to theoretically simulate them. Among all
the publicly available literature, Aspen Plus is commonly employed to model power cycles
for analysis and comparison. It is developed and licensed by Aspen Tech and is a
software for simulating commercial sequential process (Scaccabarozzi et al., 2016).

Due to the nature of my research requirements and the constraints of my professional


background, modeling the PGSs using Aspen Plus will be accommodated by finding
suitable studies. Assumptions, parameters, and evaluations based on the chosen studies
are demonstrated and discussed in Chapter 4 to 6. Furthermore, Alberta climatic and
industrial conditions will be considered in evaluating and exploring the application
feasibility of direct sCO2 power cycle technology.

24
3.1 Technology Definitions

3.1.1 NGCC Systems

A large-scale integrating research project was completed through industry collaboration


among organizations in the energy industry in Europe to define best practices and
guidelines for carbon capture research and development projects (European
Benchmarking Task Force, 2011). A similar research project was also conducted in the
US. by NETL. In these two large-scale collaborative integrating research projects, NGCC
power plant w/CC and w/o CC are evaluated and compared by a variety of metrics such
as net plant efficiency, avoided emission, and economic costs using Aspen Plus modeling
(European Benchmarking Task Force, 2011; National Energy Technology Laboratory,
2015). For the NGCC w/CC cases in both projects, aqueous amine technology was
selected.

Table 2. Overview of NGCC w/o CC cases from studies of EBTF (2011) & NETL (2015)
and the NGCC power plant of SEC w/o CC.

The NGCC w/o The NGCC w/o


Attribute NGCC of SEC CC case of CC case of
EBTF (2011) NETL (2015)
Plant Capacity
800 830 641
(MW)
2GT, 2HRSGs 2GT, 2HRSGs
2GT, 2HRSGs
Plant Configuration 1ST (multi- 1ST (multi-shaft
1ST (multi-shaft 2x2x1)
shaft 2x2x1) 2x2x1)
Rated Turbine
240/360 272 /293 211/230
Capacity (MW)
TIT (°C) 1600* 1360 1359
TOT (°C) 600 608 603
HHV Net Eff. (%) — 52.7 51.5
LHV Net Eff. (%) 60.0 58.3 57.0
(Power technology, n.d.
Source Mitsubishi Hitachi Power EBTF (2011) NETL (2015)
Systems Global, n.d.)
(Note: * means the number reported is based on personal communication without any
further confirmation from actual data.)

25
As shown in Table 2, the NGCC w/o CC system model defined by EBTF (2011) has a
capacity of 830 MW with similar advanced gas and steam turbines and a net plant
efficiency of 58.3% (LHV base), which is closer to the performance of the NGCC power
plant of SEC. On the other hand, the supplied natural gas composition in both projects
are different considering the geological, technological and market demand differences.
Table 3 indicates that the natural gas composition used in the project of NETL (2015) is
closer to the reported (Alberta Energy Regulator, 2018) natural gas composition without
impurity in Alberta. What’s more, since the assumptions made in the project of NETL
(2015) were based on North Amercian conditions, the project is a better representation
of NGCC performance in the energy industry in North America.

Table 3. Natural gas composition from projects of EBTF (2011) & NETL (2015) against
Alberta’s natural gas composition reported by AER in 2018.

NG Composition (V%) AER (2018) EBTF (2011) NETL (2015)


Methane 92.0 89.0 93.1
Ethane 5.0 7.0 3.2
Propane 1.0 0.7
Butane 0.11 0.4
3.0
Carbon Dioxide 2.0 1.0
Nitrogen 0.89 1.6

Due to the issue of confidentiality, information and data on the actual performance of
NGCC power plant of SEC is unavailable. However, as discussed above, the NGCC w/o
CC and w/CC power plant cases of NETL (2015) are used as the first two PGSs for my
research.

3.1.1.1 Power Generation System 1

PGS1 is the NGCC power plant of SEC w/o CC which is best represented by the NGCC
w/o CC power plant case from NETL (2015), and it is considered as the baseline PGS in
my research.

The following figure demonstrates the NGCC has two identical large-scale industrial gas
turbines, “F-class”, with a total rated capacity of 422 MW, and one steam turbine of 230
MW. Both the gas turbines have its own separated HRSG to feed the steam turbine
26
(European Benchmarking Task Force, 2011; National Energy Technology Laboratory,
2015). The HRSG generates three different pressure levels of steam, they are high,
intermediate, and low pressure (HP, IP, LP), to power the steam turbine and to preheat
the water used for the steam cycle leaving the condenser (European Benchmarking Task
Force, 2011; Subramanian, et al., 2017; Thattai, et al., 2014; White & Weiland, 2018;
National Energy Technology Laboratory, 2015). The three steam streams are generated
by six economizers, three evaporators, and three superheaters and one reheater of
HRSG (Thattai, Wittebrood, Woudstra, Geerlings, & Aravind, 2014). Feed water exiting
HRSGs is condensed and pumped back into the steam cycles, and the flue gas is treated
with ammonia before releasing to the atmosphere in the stack.

Figure 13. Demonstration of flow process of the NGCC power plant of SEC w/o CC.

(European Benchmarking Task Force, 2011; National Energy Technology Laboratory,


2015)

3.1.1.2 Power Generation System 2

The only difference between PGS1 and PGS2 is the additions of an Amine Unit and
carbon processing unit (CPU) to capture and compress CO2 respectively. PGS2 is the

27
most likely scenario for PGS1 with the function of capturing and compressing CO2. Figure
14 shows the flow process of the defined PGS2.

This power system has a post-combustion CO2 capture which essentially is MEA
chemical absorption (European Benchmarking Task Force, 2011; National Energy
Technology Laboratory, 2015). In general, after flue gas is treated with ammonia and CO2
is absorbed followed by amine solvent stripping and reclaiming in the Amine Unit, the
remaining flue gas is released to the atmosphere in stack, while CO2 is released for the
next process where CO2 purification, dehydration, and compression to supercritical state
take place in the CPU (National Energy Technology Laboratory, 2015). To release CO2
away from amine, additional heat is required and it is supported by the steam extracted
from LP steam turbine as Figure 14 indicates ( European Benchmarking Task Force,
2011). More detailed information about CO2 absorber, amine regeneration and
purification, and CO2 drying and compression are provided in the study of NETL (2015).

Figure 14. Demonstration of flow process of the NGCC power plant of SEC w/CC.

(European Benchmarking Task Force, 2011; National Energy Technology Laboratory,


2015)
28
3.1.2 Direct Natural Gas-fired Oxyfuel sCO2 Power Plant

Several studies on thermodynamic performance and cycle optimization of a direct sCO2


power cycle known as Allam cycle have been performed (Penkuhn & Tsatsaronis, 2016;
Scaccabarozzi et al., 2016, 2017; White & Weiland, 2018) with the property method of
Peng-Robinson (Penkuhn & Tsatsaronis, 2016; Scaccabarozzi et al., 2016, 2017).
Limited studies exist to outline a comparison of net plant electric efficiency and economic
costs between NGCC w/CC technology using aqueous amine and a direct natural gas-
fired oxyfuel sCO2 power plant. However, this specific comparison has been carried out
by a recent study of white and Weiland (2018) using the Aspen Plus software and the
Lee-Kesler-Plöcker property method. Aspen Plus was employed to evaluate
thermodynamic performance and analyze cycle optimization of the direct sCO2 power
cycle in these studies.

In a direct oxyfuel sCO2 power cycle, the working fluid stream is mostly a binary mixture
of CO2 and H2O, which is the opposite of indirect cycle, due to its direct contact with a
fuel source and oxygen which results in impurity within the working fluid (White & Weiland,
2017). Other possible binary mixtures are CO2 - O2, CO2 - N2, and CO2 – Ar (argon)
(Scaccabarozzi et al., 2016). To simulate a binary working fluid in Aspen Plus, there are
several Equation of State (EOS) to fulfill the purpose, such as Peng-Robinson, Benedict-
Webb-Rubin-Lee-Starling, and Lee-Kesler-Plöcker (Scaccabarozzi et al., 2016; White &
Weiland, 2017). The direct oxyfuel sCO2 power cycle modeled using Aspen Plus in the
study of Scaccabarozzi et al. (2016) employed Peng-Robinson EOS because it shows
slightly higher accuracy in all possible binary mixtures among the other chosen property
methods against experimental data. In contrast, White & Weiland (2017) assessed all
available property methods that can be used for modeling direct sCO2 power cycle in
Aspen Plus against REFPROP, which is the best property method used for indirect sCO2
power cycles under the binary mixture of CO2 - H2O, and the study concluded that Lee-
Kesler-Plöcker is the best property method. Both studies have their own deficiency, as
White & Weiland (2017) considered only one binary mixture where Scaccabarozzi et al.
(2016) considered all possible binary mixtures, though limited to certain property methods.

29
On the other hand, the comparison of NGCC w/CC and direct natural gas-fired oxyfuel
sCO2 power plants in the study of White & Weiland (2018) adopted the NGCC w/CC
power plant case from the research project of NETL (2015) while using Lee-Kesler-
Plöcker EOS to compare the two technologies at the same scale. Moreover, the sCO2
power cycle that was developed and modified, is similar to studies of thermodynamic
performance and cycle optimization of Allam cycle previously performed by other
researchers, but mostly based on earlier relevant works compiled by the NETL for
achieving high plant electric efficiency and purity and capture fraction of CO2. Thus, the
developed direct natural gas-fired oxyfuel sCO2 power plant of White & Weiland (2018) is
defined as PGS3 in my research.

3.1.2.1 Power Generation System 3

PGS3 is the direct heated sCO2 power plant with natural gas as the fuel source. Oxygen
can be supplied by a nearby cryogenic air separation plant. However, a cryogenic ASU
is integrated to provide high purity oxygen. Figure 15 is the flow diagram of the defined
PGS3.

The level of the plant layout complexity is lower than the previous two as it requires a
smaller number of components within the power system. In general, the flow process of
this power plant is like the one described in Chapter 2 regarding the Allam cycle. However,
there are two important features. The first is that, near 100% pure oxygen is provided by
a cryogenic ASU and mixed with one portion of the recycled CO2 stream as a safety
practice, recompressed to supercritical state (Penkuhn & Tsatsaronis, 2016;
Scaccabarozzi et al., 2016; White & Weiland, 2018) and preheated in the HTR before
entering the combustor (White & Weiland, 2018) to reduce fuel input (Scaccabarozzi et
al., 2016); the other portion of the recycled CO2 stream goes through a train of heat
recuperation which includes LTR, intermediate temperature recuperator (ITR), and HTR
for the betterment of thermal pinch points management and working fluid preheating
maximization, and in turn, to increase overall cycle and process efficiency (White &
Weiland, 2018). The second important feature of this plant is having the thermal
integration of ASU compression to provide additional heating to recycled CO2 stream in

30
HTR (as shown in Figure 15 by a thermal loop) that can’t be accomplished by the exhaust
working fluid exiting the sCO2 turbine due to the specific heat capacity and pressure
imbalance of CO2 (as explained in Chapter 2) (Penkuhn & Tsatsaronis, 2016;
Scaccabarozzi et al., 2016; White & Weiland, 2018). Meanwhile, part of this thermal
integration is also used for turbine blade cooling (White & Weiland, 2018). Note that the
CPU of PGS3 is different than that of PGS2. 5% of recycled CO2 is purged out to the CPU
for further processing to meet pipeline standards for impurity concentrations and become
a “pipeline-ready” CO2 product (Allam et al., 2017; Weiland, Shelton, Shultz, White, &
Gray, 2017; White & Weiland, 2018).

Figure 15. Flow diagram of the direct natural gas-fired oxyfuel sCO2 power plant.

(Scaccabarozzi et al., 2016; White & Weiland, 2018)

31
3.2 Comparison Scenario

As addressed in the Research Intention section in Chapter 1, to evaluate whether a direct


natural gas-fired oxyfuel sCO2 power plant outperforms a state-of-the-art NGCC w/CC
power plant, PGS2 and PGS3 are selected for comparison given that both offer the same
functionality — capturing CO2. However, when comparing with PGS1, it is important to
acknowledge the existence of an energy penalty associated with capturing and
compressing CO2 in PGS2 which, in turn, impacts on the net plant electric efficiency. A
direct natural gas-fired sCO2 power cycle is a CO2 capture ready cycle where energy
penalty does not apply to PGS3 (S. Macadam, email communication, June 06, 2018).
Thus, the issue of energy penalty is not addressed in my research for both power plants.

The comparison of these two power plants is the only comparison scenario established
for my research. General assumptions and parameters for both power plants are
referenced mainly by the studies of NETL (2015), White et al. (2017) and White & Weiland
(2018) and supported by other relevant studies. Meanwhile, assumptions and parameters
for each analytical pillar are addressed and the results of key performance indicators
(KPIs) are evaluated and compared in respective chapters (from Chapter 4 to 6). All main
sections of each power plant and the respective major components are listed in Appendix
A.

The recuperation of sCO2 power cycle can be a dynamic and complex process due to the
changing pressure and temperature between the hot and cold recycle streams within the
heat exchangers, which will result in non-linear profile changes before reaching a new
equilibrium state (Allam et al., 2017). To avoid the dynamic complexity of recuperation
during events such as power plant start up, shut down, load changing, emergency
operation, and components maintenance, steady-state simulation have been chosen and
performed for both power plants in Aspen Plus (National Energy Technology Laboratory,
2015; White & Weiland, 2018). A steady-state simulation applies when all dynamic
transients within a described system have reached their equilibrium levels where the
variation of time is assumed to be infinite.

32
There are detailed assumptions and parameters addressed and can be found in the
project of NETL (2015) on NGCC w/CC power plant case and study of White et al. (2017)
and White & Weiland (2018) on direct oxyfuel sCO2 power plant. However, only selected
assumptions and parameters are presented in this section.

3.2.1 General Assumptions and Parameters

The study scope for both PGSs includes all necessary aspects of the power plant facility
with the “fence line” (National Energy Technology Laboratory, 2015; White & Weiland,
2018), and aspects associated with CO2 transportation and storage is replaced with a
pipeline system.

Both the power plants operating under the same environmental conditions and the used
natural gas share the same composition (see Appendix B). It is assumed that both PGSs
are enclosed in a building (Nsameational Energy Technology Laboratory, 2015; European
Benchmarking Task Force, 2011; White & Weiland, 2018) in a generic plant site —
Midwestern US. (National Energy Technology Laboratory, 2015; White & Weiland, 2018),
and the capacity factor (CF) is 85% for both power plants (National Energy Technology
Laboratory, 2015; European Benchmarking Task Force, 2011; White & Weiland, 2018).
The A complete fuel-to-energy conversion and no heat loss are assumed (National
Energy Technology Laboratory, 2015; European Benchmarking Task Force, 2011; White
& Weiland, 2018). Also, both power plants use wet a cooling system and the cooling water
temperature is 16°C (National Energy Technology Laboratory, 2015; White & Weiland,
2018).

33
3.3 Objectives

Specific aspect for each analytical pillar is listed below.

• Energy – Net Plant efficiency. The net plant efficiency of a power plant is
determined by its configuration and operational practices. Some major auxiliaries’
loads are inevitable and have a tremendous impact on the net plant efficiency.
• Environmental – Raw water usage and CO2 emission. The amount of raw water
withdrawal and consumption are influenced by the designed cooling system and
its efficiency. Influencers, such as with or without exhaust gas recirculation
(Subramanian et al., 2017), impurity concentration (Scaccabarozzi et al., 2016),
and energy input for CO2 separation (Subramanian et al., 2017), could affect the
amount of avoided CO2 emission and the level of purity.
• Economic – Cost of Energy (COE). The most significant proponent concerns are
costs associated with a power plant’s life cycle-from planning to decommissioning-
and involves various aspects, such as capital cost, owner’s cost, fixed and variable
operation and maintenance (O&M) costs, and finally the COE.

Therefore, the objectives of my research are 1) to evaluate and determine whether the
direct natural gas-fired oxyfuel sCO2 power plant performs better than the NGCC w/CC
power plant (aqueous amine technology) in terms of the three analytical pillars under
assumed assumptions and parameters, 2) to evaluate whether the direct natural gas-fired
oxyfuel sCO2 technology is technologically and economically feasible with the
consideration of Alberta conditions, and 3) to explore the potential applications of the
sCO2 technology in the power generation sector in Alberta.

34
Chapter 4. Comparison – Energy Performance

4.1 Assumptions and Parameters

Natural gas is assumed to be delivered in a condition of 30 bar and 38°C and ready to
use without any additional post-fuel processing units (National Energy Technology
Laboratory, 2015) for PGS2. For natural gas used in PGS3, fuel compression is needed
to meet pressure requirement (300 bar) before combustion. The feed flow rate of natural
gas is fixed at 84,134 kg/hr for both power plants (National Energy Technology Laboratory,
2015; White & Weiland, 2018).

4.1.1 PGS2

The NGCC power cycle of PGS2 is the same as PGS1—for which plant capacity, plant
configuration and turbine capacity information are addressed in Table 2. The main
parameters of the PGS2 power plant relevant to this chapter are listed in Appendix C.
The methodology on determining these parameters are detailed in the project report of
NETL (2015).

4.1.2 PGS3

Oxy-combustor, sCO2 turbine, recuperators, and CO2 recycle compressor are the four
important components for PGS3 (Allam et al., 2017; Penkuhn & Tsatsaronis, 2016;
Scaccabarozzi et al., 2016, 2017; White & Weiland, 2018). Assumptions and parameters
for these components share some similarities but also differences among studies. The
White & Weiland (2018) study informs the brief descriptions of the four main components
,which are based on the study of Weiland et al. (2017), while some points are also
supported by other studies. Appendix D is the summary of all main parameters. Further
information regarding these four main components is available in White et al. (2017) and
White & Weiland (2018).

Oxy-combustor is modeled under the assumption of complete fuel-to-energy conversion


with a comprehensive combustion reaction designed within the capability of Aspen Plus.
Oxygen purity is designed to be 99.5%, using a low pressure cryogenic ASU, to minimize

35
impurity level (Weiland et al., 2017). The calculation of excess oxygen value is based on
the chemical reaction of complete combustion.

The sCO2 turbine handles real gas mixtures unlike a standard gas turbine that only
handles an ideal gas mixture species (Scaccabarozzi et al., 2016). The TIT is
parameterized at 1204°C which is higher than in other studies, and the required cooling
temperature for this level of TIT is modified to 400°C with the consideration of two
parameters—Gross Cooling Effectiveness and Heating Loading Parameter. The
correlation between these two parameters plus previous industrial empirical analysis
determines the mass flow rate of turbine coolant and pressure losses (Scaccabarozzi et
al., 2016). Hence, the turbine model designed is a continuous cooled expansion model,
see Figure 16, with two cooled expansion stages and one adiabatic expansion stage
(Scaccabarozzi et al., 2016; Allam et al., 2017).

Figure 16. The designed model of sCO2 turbine by General Electric Global Research.

(Limer, 2016)

Multi-streams with different thermodynamic properties flow across the recuperators,


resulting in a large variation in specific heat capacity among streams (cold and hot ends)
and water condensation (dew-point) in the exhaust stream (Scaccabarozzi et al., 2016,
2017). The recuperators are modeled to be an adiabatic countercurrent heat exchanger
and have three pinch-points. The adiabatic countercurrent design is also applied to CO2
coolers.

A series of adiabatic compressors with aftercoolers and intercoolers is applied to benefit


the integration of heat from ASU and the power cycle. The CO 2 MC is an intercooled
36
compressor that has four compression stages and three intercoolers. All compressors are
assumed to be adiabatic and have an isentropic efficiency of 85% and mechanical
efficiency of 98%.

4.2 Evaluation & Discussion

The KPIs for energy performance evaluation are Total Net Power and Net Plant Efficiency
(LHV). Detailed results and evaluations on thermal dynamic analysis and heat balance of
PGS2 and PGS3 are available in the studies of NETL (2015) and White & Weiland (2018)
supplemented by Weiland et al. (2017), respectively.

Table 4. Power summary of the two power plants.

Main Attribute (MW) PGS2 PGS3


Combustion Turbine 428 949
Steam Turbine 182 —
CO2 MC — -88
CO2 Pumps — -50
CO2 RC — -56
Generator Loss -9 -16
Total Gross Power 601 738
Auxiliary Attribute (MW) PGS2 PGS3
Feedwater Pumps -3.6 —
CO2 Capture/Removal Auxiliaries -13 —
CPU/CO2 Compression -15 -8.7
Cryogenic ASU — -116.4
(ASU compressor) — (-74.8)
(sCO2 Oxygen Compressor) — (-41.6)
Fuel Compressor — -12.7
Circulating Water Pumps -4.3 -4.2
Cooling Tower Fans -2.2 -2.2
Transformer Losses -1.8 -2.6
Others -1.8 -1.2
Total Auxiliaries’ Load -42 -148
Total Net Power 559 590

(National Energy Technology Laboratory, 2015; White & Weiland, 2018)

37
Table 4 lists the electric power output summary of major components for the two power
plants. Thanks to the nature of sCO2 turbine, PGS3 generates almost 1.6 times more
energy than that of the PGS2 combustion and steam turbines combined. However,
components of recompression and recuperation (CO2 MC, Pumps, and RC) within the
system consume about 20% (194 MW) of the power generated from the sCO 2 turbine.

The total auxiliaries’ load for PGS3 (20% of the total gross power) is about 3.5 times
higher than that of PGS2 (7% of the total gross power). Specifically, within the PGS2,
each unit of energy required by CO2 capture auxiliaries and CPU/CO2 compressor is
much higher than other components which occupy 67% of the total auxiliaries’ load (28
MW out of 42 MW). PGS3 in contrast, when considering the oxygen production of the
ASU and fuel compressor associated with meeting the operating pressure condition of
300 bar, is about 87% of the total auxiliaries’ load (129 MW out of 148 MW).

Table 5. Energy performance summary for the two simulated power plants.

Attribute PGS2 PGS3


Natural Gas Feed Flow (kg/hr) 84,134 84,134
HHV Thermal Input (MW) 1,223.0 1,223.0
LHV Thermal Input (MW) 1,105.2 1,105.2
Total Net Power (MW) 559 590
HHV Combustion/sCO2 Turbine Eff. 34.50% 58.60%
HHV Net Plant Eff. 45.70% 48.20%
LHV Combustion/sCO2 Turbine Eff. 38.10% 66.80%
LHV Net Plant Eff. 50.60% 53.40%

(National Energy Technology Laboratory, 2015; White & Weiland, 2018)

Table 5 summarizes the overall energy performance of the power plants. With the same
amount of thermal energy input, the direct natural gas-fired oxyfuel sCO2 power plant
generates more energy by 31 MW and has a higher net plant efficiency by about 3%,
despite the significant total auxiliaries’ load demonstrated by the PGS3.

The result of this evaluation completed by White & Weiland (2018) corresponds to other
studies, although the sCO2 power plant configurations are varied among studies. For

38
instance, the energy performance of the NGCC w/CC power plant case of EBTF (2011)
concluded that the LHV net plant efficiency is 49.9%; and the studies of Scaccabarozzi
et al. (2016; 2017) indicates the evaluated direct natural gas-fired oxyfuel sCO2 power
cycle has a LHV net plant efficiency of around 55%. These studies all show that the sCO2
power plant demonstrates a higher net plant efficiency and net electric power output than
a state-of-the-art NGCC w/CC power plant.

39
Chapter 5. Comparison – Environmental Performance

5.1 Assumptions and Parameters

5.1.1 Raw Water Usage

The makeup of raw water for both PGS2 and PGS3 is 50% from groundwater and 50%
from a municipal water treatment facility, and available water from the internal recycle is
used to offset overall water withdrawal (National Energy Technology Laboratory, 2015).
The assumed cooling water approach temperature in the cooling system is 5°C (National
Energy Technology Laboratory, 2015).

For a power plant, there are two water usage activities observed within the system. One
is water withdrawal, which is water removed from defined sources for use in the plant,
with the remaining water is returned to the sources. The other one is water consumption,
where withdrawn and used water does not return to the sources but rather evaporates
during the production process and ultimately is incorporated into the final products. It is
important to balance the totaling raw water demand and consumption when comparing
raw water usage between the two power plants. For both power plants, it is assumed that
the total of internal water recycle and raw water withdrawal are accounted for the total
water demand for a power plant, and the difference between water withdrawal and
consumption is the total of water returned to the sources (National Energy Technology
Laboratory, 2015; White & Weiland, 2018; European Benchmarking Task Force, 2011).

5.1.2 CO2 Emission

In the integrating collaborative research project of NETL (2015), it has been stated that
the NGCC w/CC case was designed with 90% capture to meet the new revised
Environmental Protection Agency requirement on air emission. For a typical direct sCO2
power plant, it can produce an exhaust stream with at least 98 mol% CO 2 (Weiland et al.,
2017).

It is important to balance the totaling carbon in and out within the power plant when
considering CO2 capture in a power plant. CO2 Emission is calculated based on net power
output (National Energy Technology Laboratory, 2015). Captured carbon fraction for the
40
NGCC power plant is determined as the total carbon in, subtracting the amount of CO2
emitted through stack, then dividing the total of carbon in. As for the direct sCO2 power
plant, it is assumed that CO2 in the air to ASU is 100% vented for PGS3, and the
calculation is the total of carbon captured over the total of carbon in from fuel feed
(Weiland et al., 2017; White & Weiland, 2018). CO2 purity level is adjudged according to
the oxy-combustor and configuration designed for the elimination of potential impurity
(White & Weiland, 2018).

5.2 Evaluation & Discussion

The KPIs for energy performance evaluation are Raw Water Consumption, CO2
Emission, and Carbon Captured Fraction. Detailed results and evaluations on carbon and
water balance of PGS2 and PGS3 are available in the studies of NETL (2015) and White
& Weiland (2018) supplemented by Weiland et al. (2017), respectively.

5.2.1 Raw Water Usage

Table 6. Parameters of water balance for both power plants.

PGS2
Raw Water Raw Water
Internal Recycle Water Discharge
Withdrawal Consumption
Parameter
Cooling Tower Cooling Tower Cooling Tower Cooling Tower
CO2 Capture CO2 Capture CO2 Capture
Recovery System System
CO2 Compression
Condenser Condenser
Recovery
Boiler Feedwater Boiler Feedwater Boiler Feedwater
Blowdown Makeup Makeup
PGS3
Raw Water Raw Water
Internal Recycle Water Discharge
Withdrawal Consumption
Parameter
CPU Condensers Cooling Tower Cooling Tower Cooling Tower
sCO2 Condensers

(Weiland et al., 2017; White & Weiland, 2018)

41
Table 6 lists all water required activities within both power plants. In PGS2, the raw water
withdrawal is supplied to the cooling tower (the main water consumer), the boiler
feedwater makeup, the CO2 capture system, and the condenser. Water used for the
recovery of CO2 capture and compression originates from the internal recycle. In contrast,
with PGS3 the only water consumer is the cooling tower, and water used in the CPU is
from the internal recycle. Because a direct oxyfuel sCO2 power plant does not produce
steam for power generation and producing “pipeline-ready” CO2 is an inherent function
as shown in Table 7, the plant demonstrates less raw water consumption by 17% with the
water consumed only utilized for the cooling tower.

Table 7. Raw water usage summary for both power plants.

Aspect PGS2 PGS3


Raw Water Withdrawal (L/min) 15,230 13,593
Returned Water (L/min) 3,780 4,137
Raw Water Consumption (L/min) 11,450 9,456
(Note: the numbers are normalized for per unit of net power output under the assumption
of the same thermal energy input of natural gas at 84,134 kg/hr for both power plants.)

(National Energy Technology Laboratory, 2015; White & Weiland, 2018)

5.2.2 CO2 Emission

Table 8. Parameters of carbon balance for both power plants.

PGS2 PGS3
Carbon In Carbon Out Carbon In Carbon Out
Parameter
Air (CO2) Stack Gas Vent Air to ASU Purge CO2 Knock-out
Natural Gas Feed CO2 Product Natural Gas Feed ASU Vent Stream
CO2 Knock-out CO2 Cooler Knock-out
CO2 Dryer Vent CO2 Product
CPU Vent
Tail Gas Vent

(Weiland et al., 2017; White & Weiland, 2018)

42
Table 8 lists all the parameters of carbon balance for the two power plants, which
illustrates the possibility of CO2 loss and escape, such as CO2 from knock-outs, CO2 dryer
vent, CPU vent, and tail gas vent (Weiland et al., 2017; White & Weiland, 2018).

The NGCC power plant has a CO2 emission of 40.2 kg/MWh (megawatt-hour) as Table
9 indicates. With the assumption that 100% of venting of CO2 in the air occurs for PGS3,
the only carbon source coming into the system is fuel. Direct sCO2 power cycle is a closed
cycle and theoretically there are no emissions, and the carbon capture fraction is close to
100%. However, due to the possibility that CO2 could be dissolved in water and escape
through venting, this makes gauging the loss of CO2 unquantifiable (S. Macadam, email
communication, June 06, 2018). And it is assumed that 1.8% of CO2 will be
unaccountable for due to the possibility of CO2 loss and escape. This issue could be
improved through material selection for high temperature and pressure operation, and the
improvement of operational practices (US. Department of Energy, 2015; Allam et al.,
2017). When carbon capture fraction is normalized to the same level of Carbon In, PGS3
shows a 9% increase. A detailed CO2 emission calculation is available in Appendix F.

Table 9. CO2 emission summary for both power plants.

Aspect PGS2 PGS3


Carbon In (kg/hr) 61,262 60,768
Carbon Out (kg/hr) 6,131 N/A*
CO2 Emission (kg/MWh) 40.2 N/A*
Carbon Capture Fraction (%) 90.0 98.2
CO2 Purity (mol%) 99.93 100.00
(Note: 1. the numbers are normalized for per unit of net power output under the
assumption of the same thermal energy input of natural gas at 84,134 kg/hr for both power
plants; 2. * means unquantifiable value due to the possibility of CO2 loss and escape
during operation.)

(National Energy Technology Laboratory, 2015; White & Weiland, 2018)

43
Chapter 6. Comparison – Economic Performance

6.1 Costs Estimating Methodology

Capital cost, O&M costs, and COE are the three aspects considered for the economic
performance comparison between the two power plants. The methodologies are reviewed
and selected information from the reviewed literature are demonstrated here. Information
regarding finance structure, discounted cash flow analysis, and owner’s costs are not
included in this Chapter, and they can be found in the studies of Weiland et al. (2017) and
NETL (2015). Both capital cost estimates for NGCC w/CC and direct oxyfuel sCO2 power
plants reported in the report of NETL (2015) and White & Weiland (2018) are expressed
in June 2011 US. dollar, and there is a high possibility that prices for individual component
technologies will deviate from the reference point.

6.1.1 Assumptions and Parameters

6.1.1.1 Capital Cost

Figure 17. Each capital cost level and its respective elements.

(National Energy Technology Laboratory, 2015; Weiland et al., 2017)


44
Capital cost in this section consists of Bare Erected Cost (BEC), Engineering,
Procurement and Construction Cost (EPCC), Total Plant Cost (TPC), Total Overnight
Cost (TOC), and Total As-Spent Cost (TASC) (National Energy Technology Laboratory,
2015; Weiland et al., 2017; Weiland et al., 2017). Figure 17 shows the elements of each
capital cost level. The first four capital cost levels are “overnight” costs and the costs are
part of the capital expenditure in the first year, and TASC is the total amount of capital
expenditure over the next three years (capital expenditure period is 3 years for a natural
gas-based power plant), including escalation and interest (National Energy Technology
Laboratory, 2015; Weiland et al., 2017).

The scope for all cost estimates made in this Chapter are based on the previously
described general assumptions, but do not include CO2 transportation costs through
pipeline. Information and data used to estimate the capital cost for both power plants
relies on vendors and best engineering design and practices for component technologies
(National Energy Technology Laboratory, 2015; Weiland et al., 2017).

As the NETL (2015) stated in their report, the basis used for the cost estimate for each
power plant is provided by Worley Parsons, and the costs are calibrated with actual field
data from vendors (Weiland et al., 2017). Besides capital cost estimates supported by
Worley Parsons, other key estimating elements are also considered, such as labor costs,
engineering and construction management costs, and the detailed information is listed in
the report of NETL (2015). Component technologies that are immature are estimated in
the same way as is used for mature technologies in terms of plant design (National Energy
Technology Laboratory, 2015; Weiland et al., 2017).

Multiple subcontracts is the approach used as the contracting strategy, and it is


anticipated that the power plant owner benefits the most despite associated cost and
technology risks (Weiland et al., 2017). 10% of the BEC is assigned as the contracting
fee (Weiland et al., 2017). A similar magnitude of estimates for immature component
technologies is applied for the two power plants to ensure consistency in project and
process contingencies and common cost categories process (National Energy
Technology Laboratory, 2015; Weiland et al., 2017). Process contingencies are included

45
in some subsystem levels (National Energy Technology Laboratory, 2015; Weiland et al.,
2017). Uncertainties associated with immature technologies and their integration issues
with mature technologies for both power plants are defined as 20% and 5% respectively
(National Energy Technology Laboratory, 2015). Meanwhile, to avoid any anomalies in
capital cost comparison and ensure the direction of the costs are consistent, a cross-
account comparison is applied to all specific items by referencing their technical
parameters (National Energy Technology Laboratory, 2015; Weiland et al., 2017).

6.1.1.2 sCO2 Plant Components Cost Estimates

As previously described in section 4.1.2, oxy-combustor and sCO2 turbine are the two
most innovative components and no cost estimate model is currently available in the
market (Weiland et al., 2017; White & Weiland, 2018). To overcome this challenge, the
cost is estimated by combining the market well-known costs of a gas turbine that has a
similar size and no compressor and an outer casing that can resist high pressure and is
similar to one used for HP steam turbine, but the result is highly uncertain (Weiland et al.,
2017; White & Weiland, 2018; Allam et al., 2017). After reviewing a set of comparable
gas turbine specifications with desired sCO2 oxy-turbine performance along with process
contingency application, the estimated cost for a reasonable mature sCO 2 oxy-turbine is
$51.5 Million in 2011 dollars (Weiland et al., 2017). Figure 18 shows an example of sCO2
oxy-turbine of Toshiba.

Figure 18. A 25 MW sCO2 oxy-turbine of Toshiba.

(Gas Turbine World, n.d.; Hardcastle, 2016)


46
Cost estimates for the three recuperators are based on their thermal duty and observed
temperature differential (White & Weiland, 2018). A basis is applied to LTR, ITR, CO2
cooler and condenser with a specific cost of $0.294/(W/K), and a separated estimate
approach is applied on the low and high temperature side of HTR, considering material
variability, with a specific cost of $0.253/(W/K) for the hot side displaying a temperature
equal or below 600°C and $1.318/(W/K) for above 600°C (Weiland et al., 2017; White &
Weiland, 2018).

On the other hand, pumps and compressors applied to the sCO2 power plant are most
likely to be the same as the ones that are commercially available for the utility-scale
indirect sCO2 recompression demonstration facility of the NETL, and their costs can be
quoted by experienced vendors (Weiland et al., 2017; White & Weiland, 2018).

6.1.1.3 O&M Costs

Fixed and variable O&M activities under operating and maintaining the power plants, such
as operating labor, material, and labor of maintenance, administrative and support labor,
consumables, fuel supply, and waste management, are considered in this Chapter. The
costs for both fixed and variable O&M are estimated over the power plant’s lifetime
(National Energy Technology Laboratory, 2015; Weiland et al., 2017; European
Benchmarking Task Force, 2011). The detailed description of each aspect is available in
the reports of NETL (2015) and Weiland et al. (2017). Selected aspects of capital and
O&M costs are listed in Appendix E as well as their respective assumed values.

47
6.1.1.4 COE Estimate

The revenue of a generator that is rewarded from the first year of operation is defined as
COE, and this revenue fluctuates as it is subjected to inflation over the operational lifetime
of the power plant. COE is estimated based on finance structure, economic assumptions,
TOC, O&M costs, CF, and net power output, and it can be expressed as the following
equation:

𝐶𝐶𝐹 𝑥 𝑇𝑂𝐶 + 𝑇𝐹𝐶 + 𝑇𝑉𝐶 + 𝐹𝐶


𝐶𝑂𝐸 =
𝐶𝐹 𝑥 𝐴𝑁𝑃𝑂
where:
• CCF is capital charge factor and defined as

𝑟 (1+𝑟)𝑛
𝐶𝐶𝐹 (𝑟, 𝑛) = (r is the discount rate, n is the amortization period)
(1+𝑟)𝑛 −1

• TFC is annual total fixed O&M costs


• TVC is annual total variable costs
• FC is annual fuel cost
• ANPO is annual net power output

6.2 Evaluation & Discussion

The KPIs for energy performance evaluation are Total Overnight Cost and COE. Detailed
cost estimates for each level of capital cost, each aspect within each level, and all of the
component technologies for PGS2 and PGS3 are available in the studies of NETL (2015)
and White & Weiland (2018) supplemented by Weiland et al. (2017), respectively.
Detailed cost estimates of selected items for each power plant are attached in the
Appendix G & H.

As Table 10 illustrates, the estimated cost for the cryogenic ASU, as a necessity
subsystem for the sCO2 power plant, is ranked as the highest, followed by the combustion
turbine & accessories. The CO2 removal & compression and combustion turbine &
accessories are the top two most costly items for the NGCC power plant. Due to the
nature of the sCO2 power cycle, the cost estimate of CO2 removal & compression is

48
significantly less (13 times), though the cost is higher for the accessory electric plant. No
costs are required for components in relation to steam cycle compared with the NGCC
power plant. Interestingly, the total cost estimate when combining the power generator
and other subsystems together, with the exception of the accessory electric plant, is
comparable to that of the NGCC power plant (White & Weiland, 2018).

Table 10. List of final cost estimates of outstanding components differential between the
two power plants.

Final Cost Estimates


PGS2 PGS3
Subsystems
($1000/yr)
Cryogenic ASU 0 342,954
CO2 Removal & Compression/CPU 378,178 28,293
Combustion Turbine & Accessories 134,931 263,596
HRSG, Ducting & Stack 50,316 0
Steam Turbine Generator 74,543 0
Accessory Electric Plant 59,813 107,393

(White & Weiland, 2018; National Energy Technology Laboratory, 2015)

In Table 11, the final cost estimates for every listed aspect of the sCO2 power plant are
higher than the NGCC, except TVC which is the opposite. The total TCV cost is
significantly lower due to less consumable cost, labor cost, and requirements on the
plant’s operation and maintenance (Energy, 2015; Zhu, 2018; Ahn, et al., 2015) in
contrast to the NGCC power plant, which consumes a significant amount of CO2 capture
solvent (Subramanian et al., 2017; White & Weiland, 2018). PGS3 has higher value in
TOC, and this result is expected considering the technology maturity level of the sCO2
oxy-turbine and the recuperators. On the other hand, the COE summary in the Table
indicates that the sCO2 power plant displays 5.2% lower value in COE than that of the
NGCC power plant over the expected plant’s lifetime, thanks to the nature of the sCO2
power cycle which allows the plant to produce more energy with the same amount of
thermal input while incurring less O&M costs. However, the final cost estimates of a sCO2
plant can vary between vendors for individual components. For instance, when higher
specific costs are employed to the LTR, ITR, CO2 cooler and condenser and HTR, the
49
final COE estimate is comparable with that of the NGCC plant (White & Weiland, 2018).
COE summary calculation of both power plants is in Appendix I.

Table 11. Final cost estimates and COE summary of selected aspects for both power
plants.

Final Cost Estimates ($1000/yr)


Aspect PGS2 PGS3
TPC 827,904 867,945
Owner's Costs 180,477 188,034
TOC 1,008,381 1,055,979
TFC 27,368 28,332
TVC 16,500 11,677
Fuel Cost 190,913 190,913
TASC 1,087,034 1,203,816
COE Summary ($/MWh)
Aspect PGS2 PGS3
Capital Cost 26.9 26.7
Fixed O&M 6.6 6.4
Variable O&M 4.0 2.7
Fuel Cost 45.9 43.5
Total 83.3 79.2

(White & Weiland, 2018; National Energy Technology Laboratory, 2015)

According to the criteria of the Advancement of Cost Engineering International AACE


Class 4 for advanced technology, capital cost estimate reported in this Chapter should
also incorporate a range of -15% to 30% uncertainty for the NGCC power plant (National
Energy Technology Laboratory, 2015; European Benchmarking Task Force, 2011), and
a range of -15% to 50% uncertainty for the sCO2 power plant due to their technological
maturity at their current development state (Weiland et al., 2017; White & Weiland, 2018).
Therefore, as shown in Table 12, the gaps between the two power plants are greater in
TOC when upper limits are applied as well as the capital cost per MWh. However, the
COE of the sCO2 power plant is still lower than the NGCC plant under the 30% uncertainty
scenario and only 1% greater under the 50% uncertainty.

50
Table 12. Uncertainty applied to the total cost estimates and COE for both power plants.

Uncertainty of Final Cost Estimates ($1000/yr)


PGS2 PGS3
Aspect
30% 30% 50%
TPC 1,076,275 1,128,329 1,301,917
TOC 1,256,752 1,316,363 1,489,951
Uncertainty of COE Summary
($/MWh)
PGS2 PGS3
Aspect
30% 30% 50%
Capital Cost 35.0 34.7 40.1
Total 91.5 87.3 92.7

(White & Weiland, 2018)

Both the NGCC w/CC and sCO2 power cycles are emerging technologies for the energy
industry, and therefore commercialization is not fully realized, and in turn, premiums
associated with integration and commercial application of the technologies are not
accurately addressed (National Energy Technology Laboratory, 2015; Weiland et al.,
2017). Meanwhile, the results of the cost estimates reported don't reflect the actual project
costs due to variations in considerations of the project and site-specifications (National
Energy Technology Laboratory, 2015). Given the fact that the sCO2 power cycle is the
latest emerging technology for power generation, it is reasonable to state that the TOC of
the sCO2 power plant is higher than the NGCC power plant. However, it is anticipated
that costs related to both technologies will be significantly reduced thanks to more efficient
and cost-effective engineering as the result of continuing research, development,
demonstration and market competition (National Energy Technology Laboratory, 2015;
US. Department of Energy, 2015; Weiland et al., 2017; European Benchmarking Task
Force, 2011).

51
Chapter 7. Alberta Considerations

The three analytical pillars demonstrated were under the assumed mid-western US.
conditions and they all show promising results that the direct oxyfuel sCO2 power plant
outperforms the NGCC w/CC in terms of the identified KPIs, even though it currently
requires higher TOC due to its relatively low technological maturity level. Under the
steady-state condition, there isn’t any significant difference in energy and environmental
performances in terms of geological location differential. Based on the previous chapters,
three main operating features of the direct oxyfuel sCO2 power cycle can be concluded
as high purity of oxygen, high gaseous fuel pressure, and high operating pressure inside
the combustor. These features are essential to the power cycle.

Apart from the steady-state simulation, when it comes to assessing the direct oxyfuel
sCO2 power cycle’s performance in actual operation such as start-up, shut-down, load
change, emergency operation, and partial-load operation, as well as full operability (Allam
et al., 2017), it is important to account external factors as much as internal. To further
assess the performance of the sCO2 technology in Alberta at a basic level, it is important
to consider aspects such as Alberta’s climatic and industrial conditions. As for economic
performance, considering location factor (e.g. labor cost differences) and longer
construction period with higher contingency costs, the capital cost is expected to be
higher than the reported cost estimate.

7.1 Energy & Environmental Performances

7.1.1 Climatic Condition

7.1.1.1 Ambient Condition

Given the cold climate in Alberta, parts of the province can reach as low as -40°C in winter.
The level of oxygen purity coming out of a cryogenic ASU coincides with energy
consumption (Chorowski & Gizicki, 2015). The higher the purity, the more energy it
consumes. Also, under the same purity level, higher pressure desired will result in higher
energy consumption (Scaccabarozzi et al., 2016). For a direct oxyfuel sCO2 power cycle,
high purity oxygen can be supplied at the desired pressure required by the power cycle

52
design (Allam et al., 2017; Scaccabarozzi et al., 2016; Weiland et al., 2017; White &
Weiland, 2018). This energy intensive process mainly refers to air compression in
achieving desired cooling power and overcoming a pressure drop in the system
(Chorowski & Gizicki, 2015). Thus, despite the extreme temperature in winter, it has a
negligible effect on the energy consumption of a cryogenic ASU for high oxygen purity
production, and in turn, no effect on net plant efficiency.

On the other hand, ambient condition in terms of ambient temperature (dry and wet bulb
temperatures) varies between seasons, and the changes have a different level of impact
on a power plant’s performance when scale as well as location of the power plant are
taken into consideration (Dr. M. Elgarni, personal communication, June 26, 2018). Unlike
a gas turbine, the turbine capacity is significantly influenced by ambient temperature,
pressure and relative humidity (US. Energy Information Administration, 2016). Operating
pressure is the significant influencer for a sCO2 oxy-combustion turbine as it is operated
at a significantly high temperature and pressure of between 260 bar to 300 bar
(Scaccabarozzi et al., 2016) with oxygen and fuel both pressurized to meet the
requirement before entering the combustor. Hence, the regional ambient temperature
differential has no impact on a direct sCO2 power plant, but rather on the accessibility of
cold water.

Most power plants in Alberta utilize a wet cooling method, and the makeup of raw water
withdrawal is a mixture of nearby waterbody source (e.g. groundwater, river) and a water
treatment facility. Cold winter condition results in lower ambient temperature than is
required as per the ISO condition to achieve a desired cooling water temperature. For
example, PGS3 has a minimum cycle temperature and pressure of 27°C (National Energy
Technology Laboratory, 2015) and 80 bar (Scaccabarozzi et al., 2016). Lower cooling
water temperature will have an effect on the minimum cycle temperature and pressure,
and essentially influence the sCO2 power plant performance in terms of energy
consumption of CO2 compression and net plant efficiency (Scaccabarozzi et al., 2016).
However, as per sensitivity analysis completed by Scaccabarozzi et al. (2016) conclude
on this matter, the cooling water temperature displays a positive correlation with energy
consumption of CO2 compression but negative with net plant efficiency. In other words, a
53
lower cooling water temperature will lead to a lower energy consumption of CO2
compression and higher net plant efficiency. Thus, having access to either a cold cooling
water source, a cooling tower, or an air cooled condenser is crucial for the sCO2 power
cycle (Scaccabarozzi et al., 2016; US. Department of Energy, 2015), and Albert’s climate
is suitable for that need.

7.1.2 Industrial Condition

7.1.2.1 Oxygen Supply in Alberta

As coal power plants are gradually retired from the energy market, the responsibility of
power generation serving base load will mainly fall onto NGCC power plants (Charles
River Associates, 2017). Although Alberta has high wind and solar potential, these two
energy sources are only used mainly for serving mid load (Canadian Gas Association,
2014) due to intermittency issues. A direct natural gas-fired oxyfuel sCO2 power plant can
provide service from base to peak load as much as NGCC but may not be as flexible,
especially when considering start-up.

Large power stations in Alberta, like the NGCC of SEC, have their own black start device
(e.g. diesel generator) to provide enough power to start the main power generator, which
can otherwise be energized by extracting electricity from the public grid. However, to start
up a direct natural gas-fired sCO2 power station, enough feed power is needed to support
cryogenic ASU operation and gaseous fuel compression, especially to the former.

On-site oxygen generation – the start-up time of reaching 100% load factor for a
conventional NGCC power plant is normally about 100 mins, and 36 mins for a state-of-
the-art NGCC that is capable of performing “fast start” (Eddington, Osmundsen, Jaswal,
Rowell, & Reinhart, 2017). Cryogenic ASU takes much longer (depends on required
oxygen purity and quantity) to reach 100% load factor (Chorowski & Gizicki, 2015). As
demonstrated in Chapter 4, the cryogenic ASU supplies 99.5% of the purity level of
oxygen and consumes about 16% of the total gross power. In this case, even though the
amount of energy consumed from the external source and the associated indirect GHG
emission could be higher due to a current significantly high emission intensity of 950

54
g(CO2-eq)/kWh has been reported on Alberta grid (Environment and Climate Change
Canada, 2017), the sCO2 power plant still remains superior regarding CO2 emission
thanks to the near zero emission during steady-state operation.

Oxygen supply through pipeline – Alberta has a strong industrial gases supply chain and
mature pipeline network thanks to several industrial gas producers (e.g. Praxair Inc.) that
supply to their customers nationally and globally. Oxygen supplied through a pipeline in
Alberta has a high purity level of greater than 99.5% (Praxair, n.d.) at a pressure of 30
bar (Allam et al., 2017). If oxygen is supplied through a pipeline from an adjacent air
separation plant to the sCO2 power plant, which is accessible in Alberta and economical
for a generator (Praxair, n.d.), the energy consumption and cost of an on-site cryogenic
ASU is eliminated and will significantly increase net power output and reduce capital cost.

7.2 Economic Performance

7.2.1 Climatic Condition

A power plant project generally has a longer plant construction period in Alberta due to
the impermissible weather conditions in the winter time. For instance, a NGCC facility
requires an average of 3 years to be in full commercial operational in the US. (National
Energy Technology Laboratory, 2015), but would require 3 to 5 years of development
time in Alberta (Alberta Electric System Operator, n.d.). And, in turn, the project is given
a higher CCF, and proponents will need to incorporate higher project and process
contingency costs to cover the possibilities of occurrence of unforeseeable events. These
factors ultimately affects capital cost (Chupka & Gregory, 2007; US. Energy Information
Administration, 2016).

On the other hand, many large and specialized utility components, such as turbines,
condensers, transformers, and recuperators, are often specially manufactured to meet
the particular requirements of a project. For instance, the oxy-combustor and sCO2
turbine for a direct sCO2 power plant are needed for special design and manufacturing
requirements (Allam et al., 2017; Scaccabarozzi et al., 2016; White & Weiland, 2018).
The capacity of the manufacturer and time required between order and delivery,

55
especially if these components are imported from overseas and undesirable weather
conditions are considered, could have a substantial impact on the costs of contingency.
Consequently, these constraints impact on the capital cost (Chupka & Gregory, 2007; US.
Energy Information Administration, 2016).

7.2.2 Industrial Condition

7.2.2.1 Regional Adjustments

The cost of building a power plant at different locations can vary significantly (US. Energy
Information Administration, 2016; Dr. M. Elgarni, personal communication, June 26, 2018).
For instance, the capital cost required for a NGCC power plant in Alberta is between
C$1500 to C$1950 per kW that has fixed and variable O&M costs of C$27/kW-yr and
C$8/MWh (Alberta Electric System Operator, 2017; 2018). In contrast, for a nominal
capacity NGCC in the US., it is between $978 to $1104 per kW with a value of about
$10.5/kW-yr and $2.8/MWh for fixed and variable O&M respectively (US. Energy
Information Administration, 2016). Meanwhile, the same concept applies to locations
within one region, and the gap could be as big as about 50% (US. Energy Information
Administration, 2016). Besides the costs of material input (e.g. commodity costs) due to
inflation, the other main influencer in this contrast is the labor cost and productivity,
because it impacts the overall construction cost in a substantial way (Chupka & Gregory,
2007; US. Energy Information Administration, 2016). In this context, labor cost includes
all the labor required activities and EPC services. Overall, regional adjustments involve,
but are not limited to, differences in seismic design, advancement of local technology,
urban and remote area challenges, owner’s cost, and associated overheads costs(US.
Energy Information Administration, 2016).

7.2.2.2 Generation Location

Alberta’s economy is highly correlated to the Oil & Gas Industry. As per the anticipation
of continuing relatively low oil price below the 2014 level made by forecasters, the energy
industry is more interested in small brownfield projects (pre-existing infrastructures, e.g.
transmission lines) over the larger greenfield projects (developed from scratch) (Alberta

56
Electric System Operator, 2017), due to the requirement of significantly lower capital
costs. Newly developed natural gas-fired generation facilities in the future are anticipated
to be set-up at brownfield sites (Alberta Electric System Operator, 2017). The capital cost
of the sCO2 power plant would be reduced through utilizing a brownfield coal site or a
previously identified energy project site (Alberta Electric System Operator, 2017).

7.2.2.3 Carbon Tax

Currently in 2018, the carbon tax for the energy industry is C$30/tonne, and it is estimated
that the carbon tax will remain the same for the next a few years (Alberta Electric System
Operator, 2018). Under the capacity and energy market system, the net revenue of a
generator is calculated as; the revenues rewarded from capacity and energy market
subtracts capital cost, fix and variable O&M costs, with carbon tax as part of the variable
costs (Alberta Electric System Operator, 2018). Hence, a higher emission will result in
lower revenues for the generator. In this case, between the sCO2 plant and the NGCC
plant, the generator would be benefited more from utilizing the former to generate
electricity.

57
Chapter 8. Potential Applications in Alberta

Alberta’s energy makeup, as stated in Chapter 1, is dominated mostly by coal, followed


by natural gas, with other contributors being hydro, wind, solar and biomass as
renewables. The sCO2 power cycle technology application has been studied on various
power generation systems and the outcomes are promising. In this Chapter, the
applications of sCO2 power cycle technology for Alberta’s coal generation sector, solar
energy sector, and biomass to energy as well as waste heat to energy in Alberta are
explored and discussed at a high-level.

8.1 Coal Generation Sector

Coal has played an irreplaceable role in the world power generation for decades, but as
the issue of global warming escalates, scientists and politicians in many countries are
sweeping out coal-fired power plants and replacing them with natural gas-fired power
plants to decrease GHG emissions from fossil fuel-based power generation. Alberta’s
Coal fleet is the largest in Canada, and prior to 2018, about 50% of power generated in
Alberta came from coal (Alberta Electric System Operator, 2017).

Federal coal regulations require coal power plants to match the carbon emission intensity
of a NGCC power plant of 420 tonnes/GWh (Alberta Electric System Operator, 2017),
and the provincial government has determined to completely phase out Alberta’s coal-
fired generation facilities by 2030. The implications have struck the coal industry hard and
the decision has pushed communities relying on coal as their main revenue source to find
other ways to adapt going forward. Without the commercialization of a promising and
economically feasible carbon capture and storage (CCS) technology, significantly high
amounts of carbon emission from coal power plants still contradicts the national interest
in emission-less technology for power generation, although research has been conducted
on advanced coal-fired power generation systems, such as Ultra Supercritical Cycle,
Integrated Gasification Combined Cycle, and Oxy-Pulverized Coal technologies (Park et
al., 2018).

58
Political factor plays a significant role in the fate of coal industry for power generation. In
the US., the new federal administration is lobbying to return the coal industry as power
generation on the condition that it has low or near zero emission. Thus, several studies
on evaluating indirect and direct sCO2 power cycles for various coal-fired power
generation systems have been conducted. The application of several indirect sCO2 power
cycles on coal-fired power plant have been evaluated by Park et al. (2018), and the study
suggests a result within a range of 6.2% to 7.4% net plant efficiency improvement and
that a 7.8% to 13.6% COE reduction can be expected (Allam et al., 2017). The NETL’s
study of energy performance comparison between an oxy-coal-fired circulating fluid bed
combustion for power generation using indirect recompression sCO2 power cycle and a
conventional oxy-fired circulating fluid bed supercritical steam Rankine cycle by Shelton,
Weiland, White, Plunkett, & Gray (2016) concludes that an indirect sCO2 power cycle
under a TIT of 760°C with reheat and main CO2 compressor intercooling can yield
significantly higher process efficiency and net plant efficiency with over 95% carbon
emission avoidance. In the other NETL’s study, the tech-economic performance of a coal
gasification power plant integrated with a direct-fired sCO2 power cycle (same
configuration as PGS3), conducted by Weiland et al. (2017) concludes that the integration
results in significantly higher net power output and net plant efficiency with carbon capture
fraction of 98.1% and 99.8% purity of CO2, and lower COE. However, the biggest concern
for generators integrating the technology is the greater capital expenditure compared to
the current cost for a conventional coal power plant.

sCO2 power cycle technology could give Alberta’s coal industry another chance by
significantly enhancing efficiency and reducing emission, if Alberta is willing to adopt the
technology in the coal industry for use as power generation.

8.2 Solar Energy Sector

The southeast areas of Alberta have the highest solar farm potentials (Energy Resolution
Services, 2016; Barrington-Leigh & Ouliaris, 2017), and as solar technology constantly
improves and is becoming more efficient and economically feasible, it has the potential
to occupy a notable portion of the energy market in the future (Barrington-Leigh & Ouliaris,

59
2017). CSP is another way of generating power using solar energy. CSP has the ability
to produce a higher amount of energy and generate energy even in cloudy days thanks
to its huge heat storage capacity (Binotti, Astolfi, Campanari, Manzolini, & Silva, 2017;
Crespi, Sánchez, Rodríguez, & Gavagnin, 2017; Stein & Buck, 2017). As a matter of fact,
the first national CSP is in Medicine Hat, Alberta. Thus, besides developing more solar
PV farms in the southeastern region, CSP could be the next attraction for solar energy
investors and more of it would be built in the upcoming years in Alberta.

The power cycle used in current commercial CSP is Rankine cycle with a sub-critical
steam turbine installed (Stein & Buck, 2017), and uses either molten salt, water/steam, or
synthetic oil as the working fluid (heat transfer fluid) (Binotti et al., 2017; Stein & Buck,
2017). However, the efficiency is constrained by several factors, such as the type of CSP
system, the characteristics of the heat transfer fluid, and the condenser pressure (Binotti
et al., 2017; Stein & Buck, 2017). Normally, a CSP steam turbine exhibits a net efficiency
of about 42% with a TIT of 550°C (Stein & Buck, 2017), and CSP has a solar to electricity
conversion efficiency range from 15 to 25% (US. Department of Energy, 2013). In addition,
capital cost is depended on the turbine capacity level, and higher capacity leads to lower
capital cost (Stein & Buck, 2017).

A solar tower with thermal energy storage based on an indirect sCO2 power cycle has
been proposed and studied as a promising replacement for increasing a CSP’s
conversion efficiency and reducing capital cost (Crespi et al., 2017; Stein & Buck, 2017).
Unlike the current R&D scale of sCO2 technology for natural gas-fired power plant, those
studies on CSP are mainly focused on theoretical analysis and experimental research (Li,
Zhu, Guo, Wang, & Tao, 2017). For instance, results from the study of Binotti et al. (2017)
indicates that an indirect sCO2 power cycle configurated with recompression and a main
compression intercooling achieves an overall efficiency of 24.5% with a TIT optimization
of 750°C. However, higher overall efficiency is achieved at the compromise of the receiver
efficiency (Binotti et al., 2017; Stein & Buck, 2017). It is suggested that thermal
components within a CSP system (receiver, storage, and turbine) shall be developed as
a whole to achieve plant operating optimization (Stein & Buck, 2017).

60
8.3 Biomass to Energy

Biomass contributes 3% of electricity generated in Alberta (Energy Resolution Services,


2016). Historically, both installed capacity and total generation have been increased
(Alberta Utility Commision, 2017). The growth is attributed to the increasing, and
particularly relied upon, under-utilized forest biomass and agricultural residues (Alberta
Government, 2010). Also, turning municipal solid waste (MSW) to energy is another form
of biomass to energy in the waste management sector that reduces landfill waste, and
anaerobic digestion, gasification, combustion or incineration are the technologies
employed. The major advantage of using biomass for power generation is that it can be
used for base load generation (National Energy Board, 2017).

Biomass to energy does not possess economic incentive as high as wind and solar due
to its quality and quantity availability and market challenges (Alberta Environment and
Parks, 2018; National Energy Board, 2017; Manente & Lazzaretto, 2014), and it is most
likely to remain close to its 2015 capacity level in the future (Alberta Electric System
Operator, 2017). Combustion and gasification are the two prevalent biomass to energy
conversion methods, but the conversion and heat transfer efficiencies of both methods
are relatively low (Manente & Lazzaretto, 2014). To maximize the available heat utilization
of a small to medium biomass to energy power plant under current capacity, studies such
as Manente & Lazzaretto (2014) evaluates an integration with a dual indirect sCO2 power
cycle and suggests that an improvement in conversion efficiency of about 10% can be
achieved. Even at the worst performance of the analyzed sCO2 power cycle, it still
demonstrates higher efficiency.

8.4 Waste Heat to Energy

Waste heat to energy is part of the Alberta’s renewable energy development strategy
alongside wind and solar (Alberta Environment and Parks, 2018). Alberta Innovates has
issued Expression of Interest last year to encourage all levels and scales of waste
conversion technology development in the province (Simet, 2017). The incentive is a
means to explore technologies that benefit small and urban rural communities, and

61
electricity generation is one of the value-added products identified by Alberta Innovates
(Simet, 2017).

Oil sands facilities in Alberta produce low-grade waste heat, between 150°C to 180°C,
and it is estimated that a total potential electricity of 300 MW to 600 MW can be generated
by utilizing this waste heat (JACOBS, 2014). Meanwhile, the waste heat of some orphan
or abandoned and suspended wells in Alberta have geothermal energy potential
(Canadian Geothermal Energy Association, n.d.; CISION, 2016; Stastny, 2017). An
indirect sCO2 power cycle is more suitable and performs better than Rankine Cycle on
recycling low-grade (below 270°C) to high-grade (up to 900°C) waste heat due to its
supercritical condition (Li et al., 2017; Marchionni, Bianchi, Tsamos, & Tassou, 2017).
Hence, an indirect sCO2 power cycle can be utilized to recover Alberta’s waste heat as a
means for generating power.

62
Chapter 9. Conclusions

9.1 Results of Research

My research conducted a comparative technological review between a state-of-the-art


NGCC w/CC power plant and a direct natural gas-fired oxyfuel sCO2 power plant in
energy, environmental and economic performances as the three analytical pillars. The
analysis review indicates that, with the same amount of fuel or thermal input, the sCO 2
plant outperforms the NGCC plant in total power output, net plant efficiency, raw water
consumption, CO2 emission reduction, and COE. As an emerging technology, the TOC
of the technology is higher than that of a commercialized conventional or advanced
NGCC w/CC technology and it is more attractive in a long-term because of its lower COE.

Additionally, my research also briefly investigated whether the unique Alberta climatic and
industrial conditions could have an influence on the simulated sCO2 power plant in terms
of the three analytic pillars. Ambient condition as the main climatic factor has been
investigated, and the result shows Alberta’s climate has no effect on the energy and
environmental performance of the sCO2 power plant but rather is suitable because of the
accessibility to a cold cooling water source which is essential for the plant. Secondly, high
purity of oxygen is a must for the sCO2 power plant. The condition of oxygen supply in
Alberta is assessed for that matter, and the result shows the energy industry can provide
high oxygen purity through pipeline. And lastly, both climatic and industrial condition have
a considerable impact on the capital cost of the sCO2 plant, such as undesirable weather,
vender’s capacity, labor cost difference, regional adjustment, and generation location.
However, these constraints affect all types of utility projects, and the implication of carbon
tax has a considerable impact on power generation technology selection.

At the end of my research, the potential applications on several industries have been
addressed, namely the coal industry, the solar energy sector (particularly CSP), the
biomass to energy industry, and the waste heat to energy sector in Alberta. Three
conclusions are drawn 1) sCO2 power cycle technology may possibly allow the Alberta
government the option to keep coal power plants online with near zero emission and
higher power output; 2) integrating sCO2 technology could improve CSP plant
63
performance and biomass to energy performance in Alberta; 3) Alberta’s oil sand facilities
could contribute more electricity to the public grid by utilizing waste heat with an indirect
sCO2 power cycle, and the economically valuable waste heat from abandoned or
suspended wells in Alberta could be facilitated by the technology.

9.2 Summary

Globally, theoretical research and development of the application of sCO 2 power cycle
technology for power generation has been ongoing for the past decade. Many studies
have shown the superior potentials of the technology in various aspects, and it is believed
to be on the forefront of the next potential era for electric power generation. The
advantages of sCO2 are identified as a low critical operating condition and high power
density, and in turn, the benefits of using a direct sCO2 power cycle are established as a
higher cycle and process efficiency, a higher power output, a lower plant and
environmental footprints, and near zero GHG emission. From a technology development
to commercialization perspective, if an evaluation spectrum is given to the technology,
the natural gas-fired oxyfuel sCO2 power cycle technology is scaled at 7 out of 10, as is
demonstrated by the fact that several independent or governmental collaborative
demonstration facilities (as mentioned in Chapter 1) have been built and operated to
provide field testing data and performance in an effort to bring the technology closer to
commercialization.

Alberta is transitioning from an energy-only market to a capacity and energy market


system and is committing to a low-carbon future by proactively incentivizing renewable
energy development and deploying natural gas-fired power plants to offset the capacity
loss of coal power plants. Developing NGCC w/CC is a valuable option in meeting future
energy demand and emission reduction target, but sCO2 technology may just perform
better and could give Alberta an upper hand, not only in meeting the energy demand, but
also significantly reducing GHG emissions in power generation.

Moreover, the global NRG COSIA Carbon XPRIZE competition could result in a market
demand for “pipeline-ready” CO2 product. Alongside existing practices, the participants
are also converting the CO2 into other valuable products, such as CO2 enhanced concrete,
64
liquid fuels, plastics and carbon fiber. A high concentration and purity of CO2 can benefit
those conversion technologies and products, which would alleviate the stress of finding
and developing a geological formation-based storage for captured CO2.

Ultimately, my research calls for Alberta’s attention to the sCO2 technology, particularly
direct natural gas-fired oxyfuel sCO2 power cycle technology for power generation, and
that the Alberta government and the energy industry should consider exploring the
possibility of adopting the technology in conjunction with investing in natural gas-fired
power plants and CCS technology to meet the 70% requirement of the energy projection
and emission reduction target by 2030. Future research will be required to conduct a
thorough economic analysis between a direct natural gas-fired oxyfuel sCO2 power plant
and a NGCC w/CC power plant under Alberta’s market conditions. If adopted by Alberta,
a dynamic simulation of the technology through field testing, in any given situation and
performance, is needed to acquire field engineering and operational instrumentation data
for plant optimization and commercialization.

65
References

Ahn, Y., Bae, S. J., Kim, M., Cho, S. K., Baik, S., Lee, J. I., & Cha, J. E. (2015). Review
of supercritical CO2 power cycle technology and current status of research and
development. Nuclear Engineering Technology Vol. 47, p647-661.

Alberta Electric System Operator. (2017, July). AESO 2017 Long-term Outlook.
Retrieved from https://www.aeso.ca/grid/forecasting/.

Alberta Electric System Operator. (n.d.). Capacity Market Transition. Retrieved June 25,
2018, from Proposed Gross Cost of New Entry & Net Cost of New Entry
Calculation Approach: https://www.aeso.ca/market/capacity-market-
transition/sam-working-groups/adequacy-and-demand-curve-determination/

Alberta Electric System Operator. (2018, January 26). Summary of Integrated Capacity
and Energy Revenue Modelling. Retrieved from
https://www.aeso.ca/assets/Uploads/Summary-of-Integrated-Capacity-and-
Energy-Revenue-Modelling.pdf

Alberta Energy Regulator. (2018, March). Natural Gas Methodology. Retrieved from
www.aer.ca/data-and-publications/statistical-reports/natural-gas-methodology.

Alberta Environment and Parks. (2018, April 18). Waste to Energy. Retrieved from
http://aep.alberta.ca/waste/waste-facilities/waste-to-energy.aspx

Alberta Government. (n.d.a). Climate Leadership Plan. Retrieved June 12, 2018, from
https://www.alberta.ca/climate-leadership-plan.aspx

Alberta Government. (2010, March). Alberta Biomass Opportunities. Retrieved from


https://open.alberta.ca/publications/9780778590378.

Alberta Government. (n.d.b). Genesee Generating Station Units 4 and 5 Project.


Retrieved April 29, 2018, from http://majorprojects.alberta.ca/details/Genesee-
Generating-Station-Units-4-and-5-Project/644

66
Alberta Government. (n.d.c). Renewable Electricity Program. Retrieved April 17, 2018,
from https://www.alberta.ca/renewable-electricity-program.aspx.

Alberta Utility Commision. (2017, January). Alberta Electric Energy Net Installed
Capacity and Energy Generation By Resource. Retrieved from
http://www.auc.ab.ca/market-oversight/Annual-Electricity-Data-
Collection/Pages/default.aspx.

Allam, R., Martion, S., Forrest, B., Fetvedt, J., Lu, X., Freed, D., . . . Manning, J. (2017).
Demonstration of the Allam Cycle: An update on the development status of a
high efficiency supercritical carbon dioxide power process employing full carbon
capture. Energy Procedia Vol. 114, p5948 - 5966.

Barrington-Leigh, C., & Ouliaris, M. (2017). The Renewable Energy Landscape in


Canada: A Spatial Analysis. Renewable and Sustainbale Energy Reviews, Vol
75, p809 - 819.

Binotti, M., Astolfi, M., Campanari, S., Manzolini, G., & Silva, P. (2017). Preliminary
assessment of sCO2 cycles for power generation in CSP solar tower plants. Applied
Energy, 204, p1007–1017.

Canadian Gas Association. (2014, December). Role of Natural Gas for Power
Generation. Retrieved from http://www.cga.ca/publication/the-role-of-natural-gas-
for-power-generation/

Canadian Geothermal Energy Association. (n.d.). Canadian National Geothermal


Database and Territorial Resource Estimate Maps: Alberta Thermal Favourability
Maps and Database. Retrieved June 28, 2018, from
https://www.cangea.ca/alberta-geothermal-favourability-maps.html

Charles River Associates. (2017, March 30). A Case Study in Capacity Market Design
and Considerations for Alberta. Retrieved from
https://www.aeso.ca/assets/Uploads/CRA-AESO-Capacity-Market-Design-
Report-03302017-P1.pdf

67
Chorowski, M., & Gizicki, W. (2015). Technical and economic aspects of oxygen
separation for oxy-fuel purposes. Archives of Thermodynamics, 36(1), p157–170.

Chupka, M., & Basheda, G. (2007). Rising Utility Construction Costs : Retrieved from
www.edisonfoundation.net/IEE/Documents/RisingUtilityConstructionCosts.pdf

CISION. (2016, November 03). A New Source of Clean, Renewable and Reliable
Electricity Comes to Alberta. Retrieved from https://www.newswire.ca/news-
releases/a-new-source-of-clean-renewable-and-reliable-electricity-comes-to-
alberta-599847511.html

Crespi, F., Gavagnin, G., Sanchez, D., & Martinez, G. S. (2017). Supercritical carbon
dioxide cycles for power generation: A review. Applied Energy, Vol 195, p152-
183.

Crespi, F., Sánchez, D., Rodríguez, J. M., & Gavagnin, G. (2017). Fundamental
Thermo-Economic Approach to Selecting sCO2 Power Cycles for CSP
Applications. Energy Procedia, 129, p963–970.

Duanetildan. (2016, March 13). Supercritical CO2 Used For Solar Battery Power System.
Retrieved from https://duanetilden.com/2016/03/13/supercritical-co2-used-for-
solar-battery-power-system/

Eddington, M., Osmundsen, M., Jaswal, I., Rowell, J., & Reinhart, B. (2017, September
03). Fast Start Combined Cycles: How Fast is Fact? Retrieved from
https://www.power-eng.com/articles/print/volume-121/issue-3/features/fast-start-
combined-cycles-how-fast-is-fast.html

Energy Education. (2015, August 26). Simple gas cycle plant. Retrieved from
http://energyeducation.ca/encyclopedia/Simple_cycle_gas_plant.

Energy Resolution Services. (2016, May 31). Assessment of Renewable Energy


Potential in Alberta - A Pilot Study. Retrieved from
http://www.energyrevolution.ca/uploads/7/4/4/6/74461475/albertarenewableener
gy_i-16002-rg_rep2_ra_2016-05-31.pdf.

68
Enmax. (2015, March 11). Alberta’s largest natural gas-fuelled electricity generation
facility now operational. Retrieved from https://www.enmax.com/news-
events/news/shepard-energy-centre-operational

Enmax. (n.d.). NRG COSIA Carbon XPRIZE. Retrieved April 26, 2018, from
https://www.enmax.com/about-us/carbon-xprize

Environment and Climate Change Canada. (2017, April 13). National Inventory Report
1990-2015: Greenhouse Gas Sources and Sinks in Canada. Retrieved from
https://www.canada.ca/en/environment-climate-change/services/climate-
change/greenhouse-gas-emissions/sources-sinks-executive-summary.html

European Benchmarking Task Force. (2011). European best practice guidelines for
assessment of CO2 capture technologies. Retrieved from
https://www.sintef.no/globalassets/project/decarbit/d-1-4-
3_euro_bp_guid_for_ass_co2_cap_tech_280211.pdf

Gas Turbine World. (n.d.). Gearing up for a new supercritical CO2 power cycle system.
Retrieved July 24, 2018, from http://www.gasturbineworld.com/gearing-up.html

Global CCS Institute. (n.d.). CCS Reuse Technologies. Retrieved Auguest 20, 2018, from
https://hub.globalccsinstitute.com/publications/accelerating-uptake-ccs-industrial-
use-captured-carbon-dioxide/1-co2-reuse-technologies

Government of Canada. (2018, Feburary 2). Progress towards Canada's greenhouse


gas emissions reduction target. Retrieved from
https://www.canada.ca/en/environment-climate-change/services/environmental-
indicators/progress-towards-canada-greenhouse-gas-emissions-reduction-
target.html

Hardcastle, J. (2016, November 7). Toshiba Ships Turbine for Net Power Supercritical
CO2 Power Plant. Retrieved from
https://www.environmentalleader.com/2016/11/toshiba-ships-turbine-for-net-
power-supercritical-co2-power-plant/

69
Hislop, J. (2016, April 15). America Energy News. Retrieved from
http://theamericanenergynews.com/energy-news/nte-energy-spend-2b-
developing-three-natural-gas-fired-combined-cycle-facilities.

Huck, P., Freund, S., Lehar, M., & Peter, M. (2016, March 29-31). Performance
comparison of supercritical CO2 versus steam bottoming cycles for gas turbine
combined cycle applications. Retrieved from
http://sco2symposium.com/www2/sco2/papers2016/SystemConcepts/092paper.p
df.

JACOBS. (2014). Energy Potential and Metrics Study — An Alberta Context Prepared
For Alberta Department of Energy. Retrieved from
https://haskayne.ucalgary.ca/files/haskayne/AlbertaDepartmentOfEnergy_Energy
PotentialandMetricsStudy_Mar14.pdf

Jarre, M., Noussan, M., & Poggio, A. (2016). Opertional analysis of natural gas
combined cycle CHP plants: Energy performance and pollutant emissions.
Applied Thermal Engineering Vol. 100, p304 - 314.

Layzell, D., Shewchuk, E., Sit, S., & Klein, M. (2016). Cogeneration options for a 33,000
BPD SAGD facility: Greenhouse gas and economic implications. CESAR
Scenarios Vol. 1, Issue 3, p1-54.

Leung, D. Y., Caramanna, G., & Maroto-Valer, M. M. (2014). An overview of current


status of carbon dioxide capture and storage technologies. Renewable and
Sustainable Energy Reviews, Vol. 39, p426-443.

Li, M. J., Zhu, H. H., Guo, J. Q., Wang, K., & Tao, W. Q. (2017). The development
technology and applications of supercritical CO2 power cycle in nuclear energy,
solar energy and other energy industries. Applied Thermal Engineering, 126,
p255–275.

Limer, E. (2016, April 11). This tiny turbine could be the next big thing in power. Retrieved
from https://www.popularmechanics.com/technology/infrastructure/a20359/ge-
minirotor-co2-powered-turbine/
70
Manente, G., & Lazzaretto, A. (2014). Innovative biomass to power conversion systems
based on cascaded supercritical CO2 Brayton cycles. Biomass and Bioenergy, 69,
p155–168.

Marchionni, M., Bianchi, G., Tsamos, K. M., & Tassou, S. A. (2017). Techno-economic
comparison of different cycle architectures for high temperature waste heat to
power conversion systems using CO2 in supercritical phase. Energy Procedia,
123, p305–312.

Mitsubishi Hitachi Power Systems Global. (n.d.). Gas Turbine M501G Series. Retrieved
August 22, 2018, from
https://www.mhps.com/products/gasturbines/lineup/m501g/

National Energy Board. (2017, June 02). Canada’s Adoption of Renewable Power
Sources – Energy Market Analysis. Retrieved from https://www.neb-
one.gc.ca/nrg/sttstc/lctrct/rprt/2017cnddptnrnwblpwr/bmss-eng.html.

National Energy Technology Laboratory. (2014, July 29). Techno-economic Evaluation


of State-of-the-Art and Advanced Post-combustion Capture Plant. Retrieved from
https://www.netl.doe.gov/File%20Library/Events/2014/2014%20NETL%20CO2%
20Capture/K-Gerdes-NETL-Techno-Economic-Evaluation.pdf

National Energy Technology Laboratory. (2015, July 6). Cost and Performance Baseline
for Fossil Energy Plants Volume 1a: Bituminous Coal (PC) and Natural Gas to
Electricity Revision 3. Retrieved from
https://www.netl.doe.gov/File%20Library/Research/Energy%20Analysis/Publicati
ons/Rev3Vol1aPC_NGCC_final.pdf

National Energy Technology Laboratory. (2016, February 17). Oxy-fuel Circulating


Fluidized Bed Combustion. Retrieved from www.nrcan.gc.ca/energy/coal/carbon-
capture-storage/4335

National Energy Technology Laboratory. (2017, September 13). Overview of


supercritical carbon dioxide based power cycles for stationary power generation.
Retrieved from http://www.orc2017.com/uploads/File/Presentations/234.pdf
71
Natural Resource Canada. (2016a, January 05). Carbon Capture and Storage.
Retrieved from https://www.nrcan.gc.ca/energy/coal/carbon-capture-
storage/4307

Natural Resource Canada. (2016b, January 05). Near-Zero Emissions Oxy-Fuel


Combustion. Retrieved from http://www.nrcan.gc.ca/energy/coal/carbon-capture-
storage/4307

Natural Resource Canada. (2018, March 02). Post-combustion CO2 Capture. Retrieved
from http://www.nrcan.gc.ca/energy/coal/carbon-capture-storage/4321

Net Power. (n.d.). Technology. Retrieved July 24, 2018, from


https://www.netpower.com/technology

Olmstead, D. E., & Ayres, M. J. (2014). Notes from a Small Market: The Energy-only
Market in Alberta. The Electricity Journal, Vol. 27 Issue 4, , p102 - 111.

Park, S. H., Kim, J. Y., Yoon, M. K., Rhim, D. R., & Yeom, C. S. (2018). Thermodynamic
and economic investigation of coal-fired power plant combined with various
supercritical CO2 Brayton power cycle. Applied Thermal Engineering, 130, p611–
623.

Penkuhn, M., & Tsatsaronis, G. (2016). Exergy Analysis of the Allam Cycle. The 5th
International Symposium - Supercritical CO2 Power Cycles, 1–18. Retrieved from
http://sco2symposium.com/www2/sco2/papers2016/OxyFuel/040paper.pdf

Power Technology. (n.d.). Shepard Energy Centre, Calgary. Retrieved Feburary 21,
2018, from https://www.power-technology.com/projects/shepard-energy-centre/.

Praxair. (n.d.). Oxygen. Retrieved June 29, 2018, from www.praxair.ca/en-


ca/gases/buy-liquid-oxygen-or-compressed-oxygen-gas

Scaccabarozzi, R., Gatti, M., & Martelli, E. (2016). Thermodynamic analysis and
numerical optimization of the NET Power oxy-combustion cycle. Applied Energy,
178, p505–526.

72
Scaccabarozzi, R., Gatti, M., & Martelli, E. (2017). Thermodynamic Optimization and
Part-load Analysis of the NET Power Cycle. Energy Procedia, 114, p551–560.

Shelton, W. W., Weiland, N., White, C., Plunkett, J., & Gray, D. (2016). Oxy-Coal-Fired
Circulating Fluid Bed Combustion with a Commercial Utility-Size Supercritical
CO2 Power Cycle. The 5th International Supercritical CO2 Power Cycles
Symposium, 1–18. Retrieved from
http://sco2symposium.com/www2/sco2/papers2016/OxyFuel/104paper.pdf

Simet, A. (2017, February 24). Alberta Innovates issues EOI for wate-to-energy
technologies. Retrieved from http://biomassmagazine.com/articles/14222/alberta-
innovates-issues-eoi-for-waste-to-energy-technologies

Stastny, R. P. (2017, October 16). How geothermal could transform Alberta’s


suspended and uneconomic oil and gas wells for power generation. Retrieved
from www.jwnenergy.com/article/2017/10/how-geothermal-could-transform-
albertas-abandoned-oil-and-gas-wells-power-generation/

Stein, W. H., & Buck, R. (2017). Advanced power cycles for concentrated solar power.
Solar Energy, 152, p91–105.

Subramanian, A. S., Jordal, K., Anantharaman, R., Hagen, B. A., & Roussanaly, S.
(2017). A comparison of post-combustion capture technologies for the NGCC.
Energy Procedia V114, p2631 - 2641.

Technology Centre Mongstad. (2010, july 20). Combat climate change through
technology - Amine technology. Retrieved from
www.tcmda.com/en/Technology/Amine-technology/

Thattai, A. T., Wittebrood, B. J., Woudstra, T., Geerlings, J. J. C., & Aravind, P. V.
(2014). Thermodynamic system studies for a natural gas combined cycle
(NGCC) plant with CO2 capture and hydrogen storage with metal hydrides.
Energy Procedia, 63, p1996–2007.

73
US. Department of Energy. (n.d.). Pilot Plan: Supercritical CO2 Power Cycles. Retrieved
Feburary 25, 2018, from https://energy.gov/under-secretary-science-and-
energy/pilot-plant-supercritical-co2-power-cycles.

US. Energy Information Administration. (2016, November). Capital Cost Estimates for
Utility Scale Electricity Generating Plants. Retrieved from
https://www.eia.gov/analysis/studies/powerplants/capitalcost/pdf/capcost_assum
ption.pdf

US. Department of Energy. (2013, August 20). Concentrating Solar Power Basics.
Retrieved from https://www.energy.gov/eere/solar/articles/concentrating-solar-
power-basics

US. Department of Energy. (2015). Quadrennial Technology Review 2015 Chapter 4:


Advancing Clean Eletric Power Technologies. Retrieved from
https://energy.gov/under-secretary-science-and-energy/quadrennial-technology-
review-2015.

Wang, Y., Zhaoa, L., Ottoa, A., & Robinius, M. (2017). A Review of Post-combustion
CO2 Capture Technologies from Coal-Fired Power Plants. Energy Procedia, 114,
p650-665.

Weiland, N., Shelton, W., Shultz, T., White, C., & Gray, D. (2017). Performance and
Cost Assessment of a Coal Gasification Power Plant Integrated with a Direct- Fired
sCO2 Brayton Cycle. Retrieved from https://www.netl.doe.gov/energy-
analyses/temp/Coal-fueled Direct sCO2 Baseline Plant Final Report 20170920.pdf

White, C., Gray, D., Plunkett, J., Shelton, W., Weiland, N., & Shultz, T. (2017). Techno-
economic evaluation of utility-scale power plants based on the indirect sCO2
Brayton cycle. Retrieved from https://www.netl.doe.gov/research/energy-
analysis/search-publications/vuedetails?id=2511

White, C., & Weiland, N. (2018). Preliminary cost and performance results for a natural
gas-fired direct sCO2 power plant. The 6th International Symposium -
Supercritical CO2 Power Cycles, 1–20. Retrieved from
74
http://www.sco2symposium.com/www2/sco2/papers2018/power-plants-
applications/083_Paper.pdf

White, C., & Weiland, N. (2017). Evaluation of property methods for modelling direct-
supercritical CO2 power cycles. Journal of Engineering for Gas Turbines and
Power, V140, No.1, p. 011701.

World Coal Association. (n.d.). Coal & Electricity. Retrieved April 26, 2018, from
https://www.worldcoal.org/coal/uses-coal/coal-electricity

Zhu, Q. (2018). Innovative power generation using supercritical CO2 cycles. Clean
Energy Vol. 1, Issue 1, P68-79.

75
Appendix A – Main Sections and Respective Major Components of PGS2 and
PGS3.

PGS2 PGS3
Sections
Components
Oxygen Supply Conventional Air Compressor Cryogenic ASU
Combustion NG Combustor Oxy-combustor
Combustion
2, F-class GT sCO2 Turbine
Turbine
Heat Exchanger HRSGs (LP, IP, HP) Recuperators (LTR, ITR, HTR)
1 ST
Steam Turbine Condenser —
Steam Bypass
Recirculating Pumps
Cooling System CO2 Cooler
Cooling Tower
CO2 MC
Fluid Recycle — CO2 RC
CO2 Pump
Stack Ammonia Treatment Unit —
CO2 Capturer
Amine Unit —
Condensate Pumps (LP, HP)
CO2 Dryer
CPU —
CO2 Compressor
Source (NETL, 2015; EBTF, 2011) (White & Weiland, 2018)

76
Appendix B – Assumed Environmental Conditions

Parameter Value Source


Barometric Pressure (MPa) 0.101
Environmental Ambient Temperature (°C) 15 (NETL, 2015; EBTF,
Conditions 2011; White & Weiland,
Wet Bulb Temperature (°C) 11
2018; Penkuhn &
Ambient Relative Humidity (%) 60 Tsatsaronis, 2016)
Cooling Water Temp. (°C) 16
Parameter Value (V%)
N2 77.32
O2 20.74
Air (NETL, 2015; White &
Ar 0.92
Compositions Weiland, 2018; Weiland
H2O 0.99
et al., 2017)
CO2 0.03
Total 100.0
Methane 93.1
Ethane 3.2
Propane 0.7
n-Butane 0.4 (NETL, 2015; White &
Natural Gas CO2 1.0 Weiland, 2018;
Composition N2 1.6 Penkuhn & Tsatsaronis,
Total 100.0 2016)
HHV (MJ/kg) 52.581
LHV (MJ/kg) 47.454
CO2 emission (g/kWh LHV) 203.2

77
Appendix C – Main Assumptions & Parameters of PGS2 for the Energy
Comparison

Sections Parameter Value


Temp. 38°C
Fuel Supply
Pressure 31 bar
TIT 1359°C
Combustion Turbine
TOT 603°C
Max temp. 603°C
Heat Exchanger
Exiting temp. 117°C
Temp. 566°C
Steam Turbine
Max. pressure 165 bar
Cooling System Min. cycle temp. 27°C
Stack Temp. 35°C
Heat exchanger Tapp 5°C
Amine Unit Regenerator Tapp 3°C
Outlet CO2 pressure 2.1 bar
Polytropic eff. 85%
Mechanical eff. 98%
Compression stages 8
CPU
Ave. stage pressure ratio 1.8
Min outlet pressure 4.4 bar
Max outlet pressure 153 bar

(National Energy Technology Laboratory, 2015)

78
Appendix D – Main Assumptions & Parameters of PGS3 for the Energy
Comparison

Section Parameter Value Source


(Penkuhn & Tsatsaronis, 2016;
O2 purity 99.50% Scaccabarozzi et al., 2016; White &
Weiland, 2018)
Excess O2 1%
Oxy- Max O2 stream temp. 300°C
(White & Weiland, 2018)
combustor O2 stream pressure 64.1 bar
Diluted O2 molar fraction 30%
(Allam et al., 2017; Scaccabarozzi et
Operating pressure 300 bar al., 2016; White & Weiland, 2018;
Penkuhn & Tsatsaronis, 2016)
TIT 1204°C
Pressure ratio 10.2 (White & Weiland, 2018)
sCO2 Turbine Blade coolant temp. 400°C
(Penkuhn & Tsatsaronis, 2016;
Blade metal temp. 860°C Scaccabarozzi et al., 2016; White &
Weiland, 2018)
TOT/Max Temp. 760°C
Recuperators Min Tapp 10°C (White & Weiland, 2018)
Pressure drop 1.4 bar
Cooler/Condenser (Scaccabarozzi et al., 2016, 2017;
26.7°C
(Min cycle temp.) White & Weiland, 2018)
CO2 cooler Min Tapp 11.1°C
(White & Weiland, 2018)
Pressure drop 1 bar
CO2 RC CO2 bypass 18.1V% (White & Weiland, 2018)
Isentropic eff. 85%
Mechanical eff. 98% (Scaccabarozzi et al., 2016, 2017;
CO2 MC Compressing stages 4 White & Weiland, 2018)
Intercooling stages 3
Isentropic efficiency 85%
Mechanical eff. 98% (Scaccabarozzi et al., 2016, 2017;
CO2 pump Compressing stages 2 White & Weiland, 2018)
Intercooling stages 0

79
Appendix E – Main Assumptions & Parameters for the Economic Comparison

Aspect Parameter Value Source


Plant Construction Period 3 (White & Weiland,
Power Plant Lifetime 30 2018; NETL, 2015;
Capital Expenditure Period (PGS2) 3 Weiland et al., 2017)
Capital Expenditure Period (PGS3) 5 (Weiland et al., 2017)
Inflation Rate (%) 3
(White & Weiland,
Amortization Period (yrs) 15
2018; NETL, 2015;
Interest Rate (%) 7
Weiland et al., 2017)
Capital Charged Factor (%) 11.1
(National Energy
Capital Costs
TASC Multiplier (PGS2, 33yrs) 1.078 Technology
Laboratory, 2015)
TASC Multiplier (PGS3, 35yrs) 1.140 (Weiland et al., 2017)
Cost Estimate Uncertainty Range (%) (NETL, 2015; EBTF,
• PGS2 -15/+30 2011)
• PGS3 -15/+50 (Weiland et al., 2017)
Contracting Fee (%of the BEC) 10
(NETL, 2015;
Process Contingencies (%) 5
Weiland et al., 2017)
Project Contingencies (%) 13
(White & Weiland,
Required Operators Per Shift 6
2018)
Taxes and Insurances (% of the TPC) 2
Fixed O&M
Operating Labor ($) 39.70
Costs (NETL, 2015;
Labor Burden (% of the labor rate) 30 Weiland et al., 2017)
Administrative & Support Labor
25
(% of the labor rate)
(White & Weiland,
Variable O&M
Fuel Cost ($/MMBtu) 6.13 2018; NETL, 2015;
Costs
Weiland et al., 2017)

80
Appendix F – CO2 Emission Calculation

A Molecular Mass of Carbon 12


B Molecular Mass of CO2 44
PGS2 PGS3
C Carbon Out (kg/hr) 6131 N/A
D Net Power Output (MW) 559 590
E CF 85%
F Time (hr/yr) 8766

𝐵
𝐶 (𝐴 ) 𝐸𝐹
Equation
𝐷𝐸𝐹

(Rao, 2018)

81
Appendix G – Detailed of Total Plant Cost Estimates for PGS2

Sheet 1.

82
Sheet 2.

(National Energy Technology Laboratory, 2015)

83
Appendix H – Selected Cost Estimates Information for sCO2 Power Plant

Sheet 1.

Sheet 2.

(White & Weiland, 2018).

84
Appendix I – COE Summary Calculation of Both Power Plants

PGS2 PGS3
G TOC $1,008,381,000 $1,055,979,000
H CCF 0.111
I TFC (/yr) $27,368,000 $28,332,000
J TVC (/yr) $16,500,000 $11,677,000
K FC (/yr) $190,913,000
Total Power Output
D 559 590
(MW)
E CF 85%
F Time (hr/yr) 8766

𝐻𝐺 + 𝐼 + 𝐽 + 𝐾
Equation for COE
𝐷𝐸𝐹

𝐻𝐺
Equation for Capital Cost
𝐷𝐸𝐹

𝐼 𝑜𝑟 𝐽 𝑜𝑟 𝐾
Equation for Others
𝐷𝐸𝐹

(Rao, 2018)

85

You might also like