You are on page 1of 108

DEGREE PROJECT IN ENERGY AND ENVIRONMENT,

SECOND CYCLE, 30 CREDITS


STOCKHOLM, SWEDEN 2020

Supercritical Carbon Dioxide


Brayton Cycle for Power
Generation
Utilizing Waste Heat in EU Industries

BJÖRN J. THORSSON

HADY R. SOLIMAN

KTH ROYAL INSTITUTE OF TECHNOLOGY


SCHOOL OF INDUSTRIAL ENGINEERING AND MANAGEMENT
TRITA ITM-EX 2020:471

www.kth.se
I

Abstract
The industrial sector accounts for approximately 30% of the global total energy
consumption and up to 50% of it is lost as waste heat. Recovering that waste
heat from industries and utilizing it as an energy source is a sustainable way of
generating electricity. Supercritical CO2 (sCO2 ) cycles can be integrated with
various heat sources including waste heat. Current literature primarily focuses
on the cycle’s performance without investigating the economics of the system.
This is mainly due to the lack of reliable cost estimates for the cycle compo-
nents. Recently developed cost scaling models have enabled performing more
accurate techno-economic studies on these systems. This enables a shift in
focus from plant efficiency to economics as a driver for commercialization
of sCO2 technology. This work aims to develop a techno-economic model
for these waste-heat-to-power systems. Based on the literature, waste heat
from different industries is calculated, showing that the four industries with
the greatest potential for waste heat recovery are cement, iron and steel, alu-
minum and gas compressor stations. Six different sCO2 cycle configurations
were developed and simulated for these four industries. The techno-economic
model optimizes for the highest Net Present Value (NPV) using an Artificial
Bee Colony algorithm. The optimization variables are the pressure levels,
split ratios, recuperators effectiveness, condenser temperature and the turbine
inlet temperature limited by the heat source. The results show a vast potential
for industries to cut down costs using this system. Out of the four industries
modeled, a waste heat recovery system in an iron and steel factory yielded the
highest NPV. Results show that the integration of sCO2 cycle in the cement
industry could help reduce their waste heat by 60%, whilst simultaneously
enabling them to cover up to 56% of their electricity demand. The payback
period for the four industries varies between 6 to 9 years. Furthermore, simple
recuperated sCO2 cycles with preheating are more economical than recom-
pression cycles. Even though recompression cycles have higher thermal effi-
ciency, they are limited by the temperature glide in the waste heat exchanger.
This analysis could help investors and engineers take more informed decisions
to increase the efficiency and economic return on investment for sCO2 cycles
and heat recovery at industrial sites. To encourage adoption of supercritical
CO2 cycles, a demo is needed along with more research for higher temperature
applications with special attention to mechanical integrity.
II

Sammanfattning
Industrisektorn står för cirka 30% av den globala totala energiförbrukningen
och upp till 50% av den går förlorad som spillvärme. Återskapa att spillvärme
från industrier och använda det som energikälla är ett hållbart sätt att produce-
ra el. Superkritiska CO2 (sCO2 ) cykler kan integreras med olika värmekällor
inklusive spillvärme. Nuvarande litteratur fokuserar främst på cykelens pre-
standa utan att undersöka systemets ekonomi. Detta beror främst på bristen
på tillförlitliga kostnadsberäkningar för cykelkomponenterna. Baserat på nyli-
gen utvecklade kostnadsskalningsmodeller är det möjligt att utföra mer exakta
teknikekonomiska studier på dessa system. Detta möjliggör en förskjutning i
fokus från cykeleffektivitet till ekonomi som drivkraft för kommersialisering
av sCO2 teknologi. Detta arbete syftar till att utveckla en teknisk ekonomisk
modell för dessa avfall-värme-till-kraftsystem. Baserat på litteraturen beräk-
nas spillvärme från olika industrier, vilket visar att de fyra industrierna med
störst potential för återvinning av spillvärme är cement, järn och stål, alumini-
um och gaskompressorstationer. Sex olika sCO2 konfigurationer utvecklades
och simulerades för dessa fyra industrier. Den teknisk-ekonomiska modellen
optimerar för det högsta Net Present Value (NPV) med hjälp av en artificiell
bi-kolonialgoritm. Optimeringsvariablerna är pressure levels, delade förhål-
landen, recuperatorseffektivitet, kondensortemperatur och turbininloppstem-
peraturen begränsad av värmekällan. Resultaten visar en stor potential för in-
dustrier att sänka kostnaderna med detta system. Av de fyra modellerna in-
dustrin gav ett återvinningssystem i en järn och stålfabrik den högsta NPV.
Resultaten visar att integrationen av sCO2 cykeln i cementindustrin kan bidra
till att minska deras spillvärme med 60%, samtidigt som de gör det möjligt
för dem att täcka upp till 56% av deras elbehov. Återbetalningsperioden för
de fyra branscherna varierar mellan 6 till 9 år. Dessutom är simple recupera-
ted sCO2 cykler med förvärmning mer ekonomiska än recompressioncykler.
Trots att recompressioncykler har högre termisk effektivitet, begränsas de av
temperaturglidningen i spillvärmeväxlaren. Denna analys kan hjälpa investe-
rare och ingenjörer att fatta mer informerade beslut för att öka effektiviteten
och ekonomiska avkastningen på investeringar för sCO2 cykler och värmeå-
tervinning på industriområden. För att uppmuntra antagandet av superkritiska
CO2 cykler krävs en demo tillsammans med mer forskning för högre tempe-
raturapplikationer med särskild uppmärksamhet på mekanisk integritet.
III

Acknowledgments
First off, we would like to thank Rafael Guédez, our supervisor. He played
a huge role in inspiring us towards this topic and pushing us forward. He is
one of the best professors at KTH, passionate about his students and leading
a transformation through inspiring the next generation. We would also like to
extend our gratefulness towards Silvia Trevisan, who gave us a lot of resources
and guided our way of thinking. We are also grateful for our friends and family,
who believed in us and encouraged us along the way.
Contents

1 Introduction 1
1.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Division of Work . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Literature Review 6
2.1 Waste Heat Recovery Potential . . . . . . . . . . . . . . . . . 6
2.2 Waste Heat Recovery Systems . . . . . . . . . . . . . . . . . 10
2.2.1 Organic Rankine Cycle . . . . . . . . . . . . . . . . . 11
2.2.2 Kalina Cycle . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Supercritical CO2 Cycle . . . . . . . . . . . . . . . . . . . . . 13
2.3.1 Cycle Configurations . . . . . . . . . . . . . . . . . . 17
2.3.2 Cycle Equipment . . . . . . . . . . . . . . . . . . . . 22
2.3.3 Emissions Trading System . . . . . . . . . . . . . . . 25

3 Industries and Waste Heat Estimation 28


3.1 Iron and Steel . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Gas Compression . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3 Cement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.4 Aluminum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4.1 Hall Heroult . . . . . . . . . . . . . . . . . . . . . . 35
3.4.2 Bayer Process . . . . . . . . . . . . . . . . . . . . . . 36
3.4.3 Anode Baking . . . . . . . . . . . . . . . . . . . . . 36
3.4.4 Furnace Chimney . . . . . . . . . . . . . . . . . . . . 36
3.5 Waste Heat Estimation Summary . . . . . . . . . . . . . . . . 39

4 Performance Model 40
4.1 System Optimization . . . . . . . . . . . . . . . . . . . . . . 40
4.2 Design Point Model . . . . . . . . . . . . . . . . . . . . . . . 44

IV
CONTENTS V

4.2.1 Turbomachinery . . . . . . . . . . . . . . . . . . . . 44
4.2.2 Heat Exchangers (HEXs) . . . . . . . . . . . . . . . . 44
4.2.3 Cycle Model . . . . . . . . . . . . . . . . . . . . . . 48
4.2.4 Fluid Properties . . . . . . . . . . . . . . . . . . . . . 51
4.2.5 Economic Model . . . . . . . . . . . . . . . . . . . . 51
4.3 Heat Pump Model . . . . . . . . . . . . . . . . . . . . . . . . 53
4.4 Packed Bed thermal Storage Model . . . . . . . . . . . . . . . 54
4.5 Key Performance Indicators . . . . . . . . . . . . . . . . . . . 55
4.5.1 Thermodynamic KPIs . . . . . . . . . . . . . . . . . 56
4.5.2 Economic KPIs . . . . . . . . . . . . . . . . . . . . . 57
4.5.3 Environmental KPI . . . . . . . . . . . . . . . . . . . 58

5 Results and Discussion 59


5.1 Different Industries . . . . . . . . . . . . . . . . . . . . . . . 59
5.1.1 Heat Pumps . . . . . . . . . . . . . . . . . . . . . . . 61
5.2 Different Cycles . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.3 Effect of Split Ratio on NPV . . . . . . . . . . . . . . . . . . 63
5.4 Effect of Turbine Inlet Temperature . . . . . . . . . . . . . . . 64
5.5 Effect of Recuperator Effectiveness . . . . . . . . . . . . . . . 65
5.6 System Costs . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.7 Sensitivity Analysis . . . . . . . . . . . . . . . . . . . . . . . 67
5.8 Model Verification . . . . . . . . . . . . . . . . . . . . . . . 68
5.8.1 Cycle Model Verification . . . . . . . . . . . . . . . . 69
5.8.2 Heat Exchanger Verification . . . . . . . . . . . . . . 69

6 Conclusion 72
List of Figures

2.1 Global Waste Heat as Estimated by Forman et al. (2016) . . . 7


2.2 EU Waste Heat as estimated by Papapetrou et al. (2018). . . . 8
2.3 Ts Diagrams of (a) ORC , (b) Kalina Cycle (Jouhara et al.
2018) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Temperature profiles of heat source with (a) pure fluid, (b)
zeotropic fluid, and (c) supercritical fluid. (Liu et al. 2019)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.5 Density of sCO2 at Different Pressures and Temperatures (Cunha
et al. 2018). . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6 Simple Recuperated Cycle. . . . . . . . . . . . . . . . . . . . 18
2.7 sCO2 Recompression Cycle . . . . . . . . . . . . . . . . . . 18
2.8 CO2 specific heat capacity variation in the recuperators (Liu
et al. 2019). . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.9 Recompression as Topping Cycle. . . . . . . . . . . . . . . . 20
2.10 Simple Recuperated as Topping Cycle. . . . . . . . . . . . . . 20
2.11 Preheating Cycle . . . . . . . . . . . . . . . . . . . . . . . . 21

3.1 Coke Production Process . . . . . . . . . . . . . . . . . . . . 29


3.2 Combined 2 Cycle for Coke Production . . . . . . . . . . . . 29
3.3 Simple sCO2 Cycle for Coke Production . . . . . . . . . . . . 30
3.4 Distribution of Gas Compressor Stations over Europe . . . . . 31
3.5 Variation of Flow Rate throughout the Year . . . . . . . . . . 32
3.6 Clinker Forming Process . . . . . . . . . . . . . . . . . . . . 33
3.7 Upstream sCO2 Cycle in Cement Industry . . . . . . . . . . . 33
3.8 Cement sCO2 Heat Pump . . . . . . . . . . . . . . . . . . . . 34
3.9 Cement sCO2 Heat Engine . . . . . . . . . . . . . . . . . . . 35
3.10 Schematic of Anode Baking . . . . . . . . . . . . . . . . . . 37
3.11 Temperatures along the Anode Baking Furnace . . . . . . . . 37
3.12 Variations along the Furnace . . . . . . . . . . . . . . . . . . 38

VI
LIST OF FIGURES VII

3.13 Aluminum Upstream sCO2 Cycle . . . . . . . . . . . . . . . . 39

4.1 sCO2 Recompression Cycle . . . . . . . . . . . . . . . . . . . 40


4.2 Optimization Model . . . . . . . . . . . . . . . . . . . . . . . 43
4.3 Iterative process logic flow for the cycle model. . . . . . . . . 49
4.4 Recuperated Storage . . . . . . . . . . . . . . . . . . . . . . 54

5.1 NPVs of the Different Industries . . . . . . . . . . . . . . . . 60


5.2 DPBs of the Different Industries . . . . . . . . . . . . . . . . 60
5.3 Electric Demand Sufficiency of Different Industries . . . . . . 61
5.4 The Effect of Difference in Prices on a Cement Heatpump NPV 62
5.5 The Effect of Difference in Prices on an Aluminum Heatpump
NPV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.6 NPVs of the Different Cycle Configurations . . . . . . . . . . 63
5.7 NPVs of Different Split Ratio in Recompression Cycle . . . . 64
5.8 Effect of TIT on NPV and Efficiency . . . . . . . . . . . . . . 65
5.9 Effect of Recuperator Effectiveness on NPV and Efficiency . . 66
5.10 CAPEX Share Among Main Plant Components . . . . . . . . 66
5.11 Share of Different Components in Powerblock Cost . . . . . . 67
5.12 Sensitivity Analysis of Change in Power Block Price . . . . . 68
5.13 Sensitivity Analysis of Change in Electricity Price . . . . . . . 68
5.14 The Influence of Number of Subsections on Error of Main
Heater Calculations . . . . . . . . . . . . . . . . . . . . . . . 70
5.15 The Influence of Number of Subsections on Error of Recuper-
ator Calculations . . . . . . . . . . . . . . . . . . . . . . . . 70
5.16 The Influence of Number of Subsections on Error of Preheater
Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

6.1 Sensitivity Analysis of Change in Compressor Cost . . . . . . 94


6.2 Sensitivity Analysis of Change in Turbine Cost . . . . . . . . 94
6.3 Sensitivity Analysis of Change in Recuperator Cost . . . . . . 95
6.4 Sensitivity Analysis of Change in Condenser Cost . . . . . . . 95
6.5 Sensitivity Analysis of Change in Discount Rate . . . . . . . . 96
6.6 Sensitivity Analysis of Change in DownTime . . . . . . . . . 96
6.7 Sensitivity Analysis of Change in Waste Heat Exchanger Price 97
6.8 Some Survey Answers as Received from the Cement Industry . 98
List of Tables

2.1 Number of Plants and Audits from the Campana et al. (2013)
Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Summary of Results from Campana et al. (2013). . . . . . . . 9
2.3 Recoverable heat in Exhausts of Various Steelmaking Processes
(McKenna & Norman 2010). . . . . . . . . . . . . . . . . . . 9

3.1 Summary of Aluminum Industry and extrapolating to Europe . 38


3.2 Summary of Waste Heat Estimations in the EU . . . . . . . . 39

4.1 Information needed for the waste heat flow . . . . . . . . . . . 42


4.2 Ranges of design variables for recompression cycle optimization 42
4.3 Operating assumptions for the sCO2 Waste Heat Recovery cy-
cles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4 Economic Assumptions Used in Calculating KPIs . . . . . . . 51
4.5 Industrial Electricity Prices in EU Countries . . . . . . . . . . 52
4.6 Thermodynamic Properties of Packed Bed Material (Tiskati-
nee et al. 2017, Tiskatine et al. 2017, Becattini et al. 2017,
Haenel et al. 2012). . . . . . . . . . . . . . . . . . . . . . . . 55
4.7 Values of the main parameters used for validation of the packed
bed model by Trevisan et al. (2019). . . . . . . . . . . . . . . 56

5.1 Model verification for cycle thermal efficiency using data re-
ported by Manente & Lazzaretto (2014). . . . . . . . . . . . . 69

6.1 Components’ Reference Prices and Exponents . . . . . . . . . 92


6.2 Storage Component Costs . . . . . . . . . . . . . . . . . . . . 92
6.3 Labor and Material Costs’ share of Component Costs (Wei-
land 2019) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

VIII
Chapter 1

Introduction

The amount of CO2 in the air has been increasing steadily with human ac-
tivity. This has led to global warming, an increase of temperatures on the
Earth. The power and industry sectors produce around 60% of global CO2
emissions (Lead n.d.). A lot of solutions have been addressing this problem,
by trying to reduce energy demand, reusing elements in a circular economy
or using renewables. However, a solution that addresses the two sources si-
multaneously is utilizing waste heat from industries to produce power. There
are different methods to generate power from waste heat such as Kalina and
Organic Rankine Cycles. However, they are limited by their low efficiencies.
A novel emerging technology is the supercritical Carbon Dioxide (sCO2 ) Cy-
cle. When using medium temperature waste heat, sCO2 is preferred due to
its compactness, cost and high thermal efficiency (Wang & Dai 2016, Santini
et al. 2016). It is a closed Brayton cycle that can achieve efficiencies, higher
than a traditional steam Rankine cycle at similar temperatures. An important
question is can sCO2 cycles be designed in such a way that it delivers higher
efficiencies at feasible costs and therefore yielding a high NPV for industrial
waste heat recovery?
The work hereby presented aims at studying this question in different in-
dustries and seeing the financial benefit. The methodology to answer the ques-
tion is to first model the sCO2 cycle based on the available waste heat, followed
by a techno-economic optimization of different parameters and systems. The
study also compares the results with other research and aforementioned tech-
nologies.
The waste heat can be used in the same industrial plant where it is pro-
duced. To which degree waste heat is recovered relies on different technical
and economic factors; heat source temperature, mass flow of the waste heat,

1
2 CHAPTER 1. INTRODUCTION

dust content, chemical composition and intermittency of the source.


There seems to be a great future for sCO2 cycles. It has mostly been inves-
tigated coupled with heat sources from nuclear, coal or solar plants. This work
aims at investigating sCO2 Cycles for waste heat recovery from EU industries.

1.1 Objectives
The purpose of this Thesis is to further develop the model of a sCO2 cycle,
in order to identify its potential role in the waste heat recovery market. This
is achieved through performing an optimization analysis to be able to find the
best set of different parameters that yields a financially attractive sCO2 cycle,
using a number of cost variables (KPIs). It is followed by a sensitivity analysis
to be able to recognize the most important factors that need further research.
The specific objectives are as follows:

• Identify best Industries and best integration schemes

• Develop a flexible model that can be applied to different industries

• Implement these models using real-life data

• Investigate the techno-economic performance of these systems and fi-


nancial feasibility

• Identify the market conditions that make these models financially attrac-
tive.

• Perform sensitivities to electricity prices and advances in technology

1.2 Limitations
This thesis is limited to optimizing an sCO2 cycle for the highest NPV in 4
different industries using 6 different configurations. It solely focuses on the
EU, and results may vary in other locations. There are some possible limita-
tions in this study. First is that the cooler component was simplified to save
computing power as is discussed in Section 4.2.2. For further studies, more
computing power will yield more accurate results. Secondly, there is a possi-
bility of error in waste heat estimations carried out by other researchers and
the authors of this paper. Another problem was that the survey sent out was
filled by a few industry players. Coupled with the fact that there is a limited
CHAPTER 1. INTRODUCTION 3

access to real data from industries. The aforementioned three limitations can
be overcome in the future by partnering with industries that are interested in
this technology. Thirdly, The cost model for the waste heat recovery heat ex-
changer is not as details as the model used for other components, which may
lead to misleading results. Development of a detailed cost model for that com-
ponent would enhance the accuracy of the economic analysis. Last, there is a
limited knowledge of sCO2 cycle component performance. However, as time
progresses and the technology evolves, experience will increase.
4 CHAPTER 1. INTRODUCTION

1.3 Methodology
1. Literature Review of the different topics presented here such as the sCO2 cy-
cle, recompression cycle, different industries such as Steel and Cement. These
are presented in Chapter 2. It consists of the following steps:

• Understanding sCO2 Cycles and components

• Gathering info about waste heat in EU

• Gathering info about major industries and how they operate

• Contacting one industry and getting real data

• Analyzing the industries and calculating waste heat. For each industry,
this is the general process:

– Check EU production of the desired industry.


– Analyze different existing factory sizes and their heat streams (flow
rate, temperature and specific heat capacity)
– Figure out which heat streams can be utilized (some are very im-
portant for the industry and some might be contaminated)
– Check their future plans and the vision for 2050
– Calculate a factor of Waste Heat per tonne of production
– Multiply Heat factor with EU production capacity
– Validate using research and existing literature review

• Researching Waste Heat Recovery Technologies

• Investigating Emission Trading Scheme (ETS)

2. Model Development and Implementation. This part is two-fold. The first


part aims at modeling waste heat recovery for the different industries. Second
part aims at developing a MATLAB model of the sCO2 cycle along with its
complex recuperators and incorporating the heat from different industries. The
Model Development is explained in Chapter 4. The process is as follows:

• With the gathered data, the industries are analyzed. Therefore it is feasi-
ble to see where the sCO2 cycle can add value to the plant owners. The
plant size to be used in the model will be an average plant size in the
EU.
CHAPTER 1. INTRODUCTION 5

• Designing the sCO2 Cycle and the parameters to use

• Modeling the sCO2 Cycle

• Verification by comparing the efficiency to other studies

3. Techno-economic optimization of different industries using the developed


model of sCO2 Cycle. The process is explained in further detail in Chapter 4.

• Gathering Costs of different components

• Profitability analysis using NPV and payback period

• Optimizing the cycle using a Metaheuristic algorithm for the highest


NPV

4. Analyzing Results. The final optimal results of each industry are analyzed.

• Calculating exergy efficiency and waste heat utilized

• Calculating CO2 mitigated

• Comparing results of different industries

• Figuring out obstacles different industries have to overcome and which


industry is most suited to start implementing the sCO2 Cycle.

1.4 Division of Work


The work load was equally divided among the two authors. Both were respon-
sible for deeply understanding the cycle and how it works. Both also worked
on the literature review and data gathering about the different industries. Both
authors worked on developing the model structure for the power cycle, i.e.,
how it would calculate all thermodynamic states through the cycle. While
Björn went into the deep details of the cycle components, optimization meth-
ods, the calculations of the heat exchangers, speeding up the code, and running
calculations. Concurrently, Hady went through all the data gathered and per-
formed the estimations of the waste heat potential for EU industries and began
writing the literature review and the report. Both authors worked on analyzing
the results once the optimization was done and then both worked on finishing
the writing of the thesis.
Chapter 2

Literature Review

This chapter contains the information gathered at the beginning of the work. It
also provides the foundation on which the modeling will be performed. In spe-
cific, this chapter will focus on the amount of unutilized industrial waste heat
in the European Union, the major industries, waste heat recovery technologies
and Emission Trading Scheme.

2.1 Waste Heat Recovery Potential


Forman et al. (2016) calculated that there are approximately 68 PWh of wasted
heat every year globally. The industrial high grade heat at temperatures above
300◦ C was estimated to be 3.4 PWh. Their methodology was checking how
much energy sources are being used (such as coal and petroleum), and how
much useful energy each sector actually consumes. The remainder is the waste
heat. This is illustrated in Figure 2.1.
Firth et al. (2019) continued building on Forman’s research and delved into
the industrial sector . They estimated the global waste heat from industries in
2030 under different scenarios. The biggest industrial potential was in non-
metallic minerals (e.g. cement and glass) and Iron & Steel industries.
Some researchers focused on low grade waste heat in the EU. For example,
Kosmadakis (2019) investigated how heat pumps can be used to upgrade low-
grade waste heat industry utilization. Cayer et al. (2010) investigated using a
transcritical CO2 (tCO2 ) cycle to generate electricity from industrial low grade
heat. However, the reported thermal efficiencies are low between 6-9%.
Hammond & Norman (2014) calculated the waste heat coming out of in-
dustries in the UK. It was based on research and data gathered by McKenna
& Norman (2010). Their main data source was the UK National Allocation

6
CHAPTER 2. LITERATURE REVIEW 7

Figure 2.1: Global Waste Heat as Estimated by Forman et al. (2016)


8 CHAPTER 2. LITERATURE REVIEW

Plan for the EU Emissions Trading Scheme, which has data of 90% of energy
intensive industries. They estimate their error in calculations to be ± 33%. Pa-
papetrou et al. (2018) took Hammond’s research and scaled it up, accounting
for different energy intensities and different energy efficiency improvements.
They showed that for the EU alone, the potential of industrial waste heat is 314
TWh/year, with 33% at temperatures of 100-200◦ C (100 TWh/year), 25% be-
tween 200-500◦ C (78 TWh/year) and the rest above 500◦ C (124 TWh/year).

Figure 2.2: EU Waste Heat as estimated by Papapetrou et al. (2018).

Figure 2.2 shows that there is a huge potential in the Iron & Steel Industry.
It is followed by non-metallic minerals (cement and glass) and then the Non-
ferrous metals (Aluminum industry).
Campana et al. (2013) focused on four main industries in the EU; Steel,
Gas Compression, Cement and Glass. They analyzed 44 audits and feasibility
studies of different factories. The number of plants modeled and audits inves-
tigated is shown in Table 2.1 and a summary of their results is shown in Table
2.2.

Table 2.1: Number of Plants and Audits from the Campana et al. (2013) Study
Number of Number of
Industry
plants considered audits analyzed
Iron and Steel 399 8
Cement 241 21
Glass 58 5
Gas Transmission 613 10
CHAPTER 2. LITERATURE REVIEW 9

Table 2.2: Summary of Results from Campana et al. (2013).


T Mass flow rates EU Energy recovery
Industry ◦
( C) (kg/s) (GWh)
Steel 200-250 3-8
5984
Steel 300-320 8-20
Cement 260-420 15-110 4592
Float Glass 300-470 20-45 628
Gas Compression 300-500 30-70
10432
Gas Storage 300-500 30-70

McKenna & Norman (2010) concluded that even though Aluminum has a
lot of waste heat, but most of it is low grade waste heat. However, they have
showed that there is high grade waste heat in the steel industry as can be seen
in Table 2.3.
McBrien et al. (2016) also confirm this. They mention the importance of
upgrading the waste heat recovery systems installed in steel mills to be able to
utilize it more.

Table 2.3: Recoverable heat in Exhausts of Various Steelmaking Processes


(McKenna & Norman 2010).
Process Exhaust Recoverable Exhaust
Process Temp. Temp. heat in exhaust Stream
◦ ◦
( C) ( C) (GJ/t steel) (gas/solid)
Coke ovens 1100 1100 0.12 - 0.24 Hot coke (s)
Cooler and
Sintering 1350 350 0.49 - 0.97
exhaust gas (g)
Blast Blast furnace
1500 150 0.16 - 0.31
furnace exhaust gas (g)
Basic Basic oxygen
oxygen 1600 1600 0.10 - 0.20 furnace gas
furnace (g)
Continuous
980 800 0.25 - 0.50 Cast slabs (s)
casting
Hot rolling 900 900 0.31 - 0.62 Steel (s)
10 CHAPTER 2. LITERATURE REVIEW

Table 2.3: Continued.


Technology for
Process
heat recovery
Coke ovens Dry quenching
Advanced
Sintering
sintering
Top-pressure
Blast
recovery turbine,
furnace
dry cleaning
Basic
Gas recovery/
oxygen
Boiler
furnace
Continuous Radiant heat
casting boilers
Water spraying,
Hot rolling
Heat pumps

Lu et al. (2016) have shown that the glass industry has the highest waste
heat to power capacity per ton compared to other industries . Johnson et al.
(2008) have showed that for the glass industry in the US, there are different
grades of unrecovered waste heat. However, Campana et al. (2013) showed
that this is not the case in the EU. In the EU, the heat is utilized better. Since
there is a small number of plants, it has a small potential, therefore it will
not be considered in this work. This work will focus on Iron & Steel, Gas
Compressor stations, Cement and Aluminum industries.

2.2 Waste Heat Recovery Systems


Waste heat recovery methods involve the capture and transfer of excess heat
from a process and using it as an extra heat source. The energy from this heat
source can be used directly as heat, either within the same process or in some
other process. Another option is to utilize the heat source to generate electrical
and mechanical power (Naik-Dhungel 2012).
There are various heat recovery systems available, which are used for the
capture and recovery of waste heat. These systems primarily consist of com-
mon heat exchangers technologies such as air preheaters, recuperators, furnace
regenerators, rotary regenerators, heat wheels, run around coils, regenerative
CHAPTER 2. LITERATURE REVIEW 11

and recuperative burners, plate heat exchangers, economisers, heat pipe heat
exchangers, waste heat boilers, and direct electrical conversion devices. These
technologies all function by the same principle of capturing, recovering and
exchanging heat with a potential energy content in a process (Jouhara et al.
2018).
The use of thermodynamic cycles enables heat recovery from waste sources
to be conducted to produce electrical energy and improve the energy efficiency
of a process (Costiuc et al. 2015). Conventionally, this has been performed us-
ing water as the working fluid in the steam Rankine cycle. However, a compar-
ative thermodynamic analysis by Nemati et al. (2017) suggested that the use of
thermodynamic cycles employing organic working fluids is a promising way
of waste heat recovery from low or medium grade heat sources. Therefore,
in this chapter these thermodynamic cycles for waste heat recovery will be
reviewed.

2.2.1 Organic Rankine Cycle


The Organic Rankine Cycle functions under the same working principle as
the conventional Steam Rankine Cycle. However, instead of using water as
the working fluid the system uses organic fluids with low boiling points and
high vapor pressures (Quoilin et al. 2013). The use of an organic fluid in a
Rankine cycle makes the system suitable for utilizing low and medium grade
waste heat along with other energy sources such as biomass, geothermal and
solar applications (Rentizelas et al. 2009, Song et al. 2020, Freeman et al.
2015).
A typical ORC system consists of a set of heat exchangers (preheater and
evaporator), turbine, recuperator, pump, and condenser. The organic fluid cy-
cles through the preheater and evaporator where it is heated by the energy
source, vaporizes and becomes superheated vapor.
The vaporised organic fluid then passes through a turbine where it expands
causing the turbine to spin and generate electricity. The vapor then enters the
recuperator to reduce its temperature and utilizing that energy to preheat the
organic fluid at a later stage in the cycle.
The fluid is cooled further in the condenser where the organic vapor con-
denses back into a liquid. Finally, the condensed organic fluid is then re-
pressurized in the pump before entering the recuperator to get heated and the
cycle restarts.
The design and performance of an ORC system is highly dependent on
the working fluid that is selected and that fluid’s thermodynamic properties
12 CHAPTER 2. LITERATURE REVIEW

along with its environmental and safety criteria (Saleh et al. 2007). A ther-
modynamic analysis conducted by Douvartzides & Karmalis (2016), which
considered 37 different working fluids, showed that the appropriate selection
of a working fluid can increase the overall plant efficiency by 5.5% and reduce
fuel consumption by 12.7%.

2.2.2 Kalina Cycle


The Kalina cycle is very similar to the ORC. However, it uses a mixture of
water and ammmonia as its working fluid (Mlcak n.d.). It has the same com-
ponents as an ORC, but with an extra recuperator and separator. The added
benefit of using a Kalina cycle is that heat exchange does not happen at con-
stant temperature during boiling, but rather there is a temperature glide as
seen in Figure 2.3. This results in a higher cycle efficiency (Rogdakis & Lolos
2015). In single fluid cycles such as ORC, the working fluid rises to the boiling
temperature and then it reaches supercritical or superheated stage. However,
in binary fluid cycles, each fluid’s temperature rises independently since each
have different boiling points. This leads to better thermal matching with the
heat exchangers as the heat/cold sources do not have satisfy one set of work-
ing fluid parameters (Martínez 1992) . It is more efficient at medium grade
temperatures above 200◦ C (Milewski & Krasucki 2018).

Figure 2.3: Ts Diagrams of (a) ORC , (b) Kalina Cycle (Jouhara et al. 2018)
CHAPTER 2. LITERATURE REVIEW 13

2.3 Supercritical CO2 Cycle


Steam cycles have been in use since the 18th century. They have reached effi-
ciencies of 40% (SWEPCO Fact Sheet n.d.) and expected to rise to 50% using
advanced ultrasupercritical cycles (Weiland & Shelton 2016).
An alternative cycle to steam which has been gaining attention over the
last years is the sCO2 cycle. First mention of sCO2 cycle is in 1948, when
Sulzer bros patented a CO2 Brayton cycle (Schmidt 1970). Since then many
countries began investigating it (Gokhshtein & Verkhivker 1969, Gokhshtein
et al. 1971) Angelino in Italy (Angelino 1968, 1969) , Feher (1968) in the
United States , Sulzer Brown – Boveri in Switzerland (Strub & Frieder 1970).
Feher (1968) proposed a cycle above the critical temperature and pressure.
He concluded that this cycle can be highly efficient and have small compo-
nent sizes. Angelino (1969) focused on condensation cycles. However, he
concluded that this cycle is more efficient than steam cycles. He found that
recompression cycles are very efficient. He also concluded that CO2 is ther-
mally stable up to 1500◦ C and 40MPa. In addition, he recognized how CO2
has lower specific volume and the importance of minimizing the effect of heat
capacity changes.
Feher (1968) designed a 150kW supercritical cycle. However, to achieve
part-time load they used a parasitic load bank, as they could not find a suitable
bypass valve.
In the following decades, different layouts were designed and modeled.
It was found that the recompression achieved the highest efficiencies without
increasing complexity (Watzel 1971, Pfost & Seitz 1971). In 1976, General
Electric compared many cycles for use in applications using coal and coal de-
rived fuel. Unfortunately, they concluded that sCO2 is very expensive. How-
ever, this was due to using way higher pressures and temperatures than needed.
None of these studies were taken further though due to the technical limi-
tations of turbomachinery and heat exchangers at that time along with the lack
of heat sources. Recently, there was a revival of interest in sCO2 again. In
1997 the Czech Technical University and in 1999 (Petr & Kolovratnik 1997,
Petr et al. 1999), 2000 in MIT in USA (Dostal et al. 2002) then in 2001 the
Tokyo Institute of Technology (Yasuyoshi et al. 2001).
Many pilot projects have been implemented to further study this cycle.
Sandia has implemented a 1MW cycle (Conboy et al. 2012, 2013, Fleming
et al. 2012). China also made some progress (Wang et al. 2014). They initially
focused on the thermotechnical calculations of the components used in the
Demo project in Czech (Vesely et al. 2014).
14 CHAPTER 2. LITERATURE REVIEW

tCO2 cycles, which operate both at subcritical and supercritical conditions


have also been investigated (Zhang et al. 2007, 2005, 2006, Zhang & Yam-
aguchi 2008, 2011, Pan et al. 2016, Ge et al. 2018, Li et al. 2018, Shi, Shu,
Tian, Huang, Chen, Li & Li 2017, Li et al. 2017, Shi, Shu, Tian, Huang,
Chang, Chen & Li 2017). However, they have lower efficiencies than sCO2
Cycles.
Support continued for sCO2 . US Department of Energy (DoE) has been
supporting it. Several companies such as Echogen, GE and Netpower along
with different research facilities (Sandia, Oak Ridge, KAIST (Korea Advanced
Institute of Science and Technology)) have all been working on developing the
sCO2 cycle. Within the EU, the EC has funded two projects (HeRo and Flex)
(sCO2-flex 2019). They aim at developing a small scale Brayton sCO2 cycle
and a modular flexible coal plant based on sCO2 cycle.
Recently General Electric, led by Vinnemier published a paper talking
about using sCO2 cycles as storage in thermal plants. Their modeling shows
that the cycle can have a round trip efficiency of 60% (Vinnemeier et al. 2016).
However, this had no economic modeling in it. Therefore it might be very ex-
pensive and not financially attractive. They also compare using different fluids
such as air and Argon and the benefits of each. They show that tCO2 is ben-
eficial for inputs of low grade temperatures. According to their study, current
reasonable design temperatures are 300-600◦ C and recuperation 0.25-0.6 (due
to technical and efficiency factors for heat pumps).
Carbon dioxide enters supercritical conditions at temperatures higher than
31.1◦ C and pressures above 74 bars. There it does not behave as an ideal gas.
Its behavior is very sensitive to the changes in pressure and temperature. sCO2
Cycles have the potential of having a higher thermal efficiency than Rankine-
Steam cycles. When used for waste heat recovery, it performs better than the
steam cycle (Kizilkan 2020). In the steam cycle, the heat exchange is limited
by the narrow range of allowable temperatures. However, the sCO2 Cycle can
recover heat at different waste temperatures and utilize it efficiently.
Its low critical temperature, makes it easier to use air as a cooling medium.
Using air instead of water will reduce the impact that the system has on the
environment (Liu et al. 2019). Furthermore, CO2 is non-flammable, currently
in excess in nature and inexpensive. It has a superior thermal stability.
sCO2 cycles achieve high efficiencies over a broad range of temperatures
from different heat sources. It is made of compact components and there-
fore smaller and cheaper than steam cycles. This is because it has a relatively
higher working pressure. It does not require water conditioning or conden-
sate control. This is necessary in steam generation to avoid corrosion, fouling
CHAPTER 2. LITERATURE REVIEW 15

and scaling. Dry CO2 is non-corrosive, non-fouling and non-scaling. It usu-


ally does not require multiple stages of energy exchange (boiler, superheater,
several turbines, heat recovery steam generator, economizer). Therefore the
system is smaller and cheaper. sCO2 cycle has a relatively shorter start-up
time. Even when operating at lower temperatures of around 204◦ C, they gen-
erate 13% more energy than an HRSG and 48% more than an ORC system
(Persichilli et al. 2011).
In the Brayton cycle, sCO2 avoids the pinch problem in the heat exchanger.
As shown in Figure 2.4, there is no pinch point for an sCO2 heating process.

Figure 2.4: Temperature profiles of heat source with (a) pure fluid, (b)
zeotropic fluid, and (c) supercritical fluid. (Liu et al. 2019)

This pinch point problem is usually present for other working fluids, such
as pure and zeotropic working fluids used in ORCs and Kalina cycles. Avoid-
ing this problem reduces the irreversibility of heat exchangers. It also allows
for more heat absorption.
When compared with Helium or other ideal gases, CO2 requires less com-
pressor work. It takes around 30% of the turbine output work rather than 45%
as is the case with Helium (Dostal et al. 2004). The properties of CO2 ap-
proach that of an incompressible fluid when operating near the critical point,
which means that it requires less work for compression (Liu et al. 2019). In
turn, one stage compressors can be used. Turbines and heat exchangers are
also more compact. It can achieve 46% thermal efficiency at 550 ◦ C, which
helium achieves at 800◦ C (Dostal et al. 2004). Compression is usually done
close to the critical point as shown in Figure 2.5, to keep the compressor power
minimum.
The turbine power is not affected by the operating pressures. The pres-
sure ratio and temperature determine the output. However, the compressor is
affected by the operating pressures (Dostal et al. 2004). This is because the
16 CHAPTER 2. LITERATURE REVIEW

Figure 2.5: Density of sCO2 at Different Pressures and Temperatures (Cunha


et al. 2018).

density of CO2 changes around the critical point. It decreases drastically after
it passes the critical point. This is why sCO2 cycles are more efficient than
ideal gas Brayton cycles. The cost of CO2 fluid is 90 % less than Helium cycle
and 98% less than the organic fluid R-134a (Liu et al. 2019).
The density is not the only property that varies with pressure and tem-
perature changes. The specific heat capacity also changes greatly. So within
the heat exchanger/recuperator, the heat exchange varies greatly, and the min-
imum temperature difference happens within the recuperator rather than the
inlet/outlet of the recuperator (Dostal et al. 2002). Therefore, it is essential to
evaluate the properties along the heat exchanger because a simplified analy-
sis will not catch that behavior. For Helium and other ideal gases, it does not
change significantly. Therefore the efficiency depends only on the temperature
and pressure ratio. However for sCO2 the operating pressures have a signif-
CHAPTER 2. LITERATURE REVIEW 17

icant impact. Therefore, this has to be analyzed to determine the optimum


cycle efficiency and recuperator size. Some disadvantages are material corro-
sion at temperatures exceeding 500◦ C and requiring high pressures to attain
high efficiency (Liu et al. 2019).
Another beneficial point of sCO2 cycles is that the start-up time is currently
below 5 minutes, which is very fast for a thermodynamic system (Mercangöz
et al. 2012).

2.3.1 Cycle Configurations


Turbine work can be increased by recovering heat from the recuperator or by
adding a reheating process. Compression work can be decreased by reducing
the compressor inlet temperature along with adding intercooling. However,
the complexity and the difficulty in controlling the system might be increased
substantially when adding additional coolers or heaters (Liu et al. 2019). Re-
searchers have carried out research on different cycle configurations. Exam-
ples are not limited to simple recuperated, condensation, precompression, re-
compression, split expansion, partial cooling, cascade, preheating, simple re-
heat, and double reheat cycles. This work uses a more detailed heat exchanger
model than other work to compare 6 different cycle configurations.

Simple Recuperated Cycle


Adding a recuperator to recover heat can increase the efficiency of the cycle.
However, the greater thermal capacity of sCO2 on the high pressure side can
lead to an internal pinch point within the recuperator, which can lower the
cycle efficiency (Liu et al. 2019). The internal pinch point happens due to the
large difference between thermal capacities on either sides of the recuperator.
The cycle schematic can be seen in Figure 2.6.

Recompression Cycle
The pinch point problem present in the simple recuperated cycle can be avoid
by using split flow cycles such as the recompression cycle. The cycle config-
uration (shown in Figure 2.7) is more complex, having two compressors and
two recuperators.
18 CHAPTER 2. LITERATURE REVIEW

Figure 2.6: Simple Recuperated Cycle.

Figure 2.7: sCO2 Recompression Cycle

The CO2 leaving the low temperature recuperator (LTR) is split into two
streams. The majority of the flow enters the cooler and is compressed by the
main compressor whereas the remaining flow is directly compressed in the
recompressor. The stream leaving the main compressor is heated in the LTR
before mixing with the stream from the recompressor and being further heated
in the high temperature recuperator (HTR) (Liu et al. 2019).
CHAPTER 2. LITERATURE REVIEW 19

As the CO2 approaches its critical point, its heat capacity increases dras-
tically. Therefore the hot CO2 after the turbine has a lower thermal capacity
than the cold CO2 on the high pressure side of the recuperator. This decreases
the cycle efficiency because the temperature the CO2 rises to is limited. The
recompression cycle helps in overcoming this obstacle (Baldwin et al. 2015).
As can be seen in Figure 2.8, the difference between thermal capacities is
much greater in the LTR, than in the HTR. Therefore, having only a fraction
of the fluid pass through the cold side of the LTR, the pinch point problem can
be avoided. This creates a better match of thermal capacities between the hot
and cold streams, giving a lower temperature pinch point in the recuperator,
which improves the heat exchange (Trevisan et al. 2019).

Figure 2.8: CO2 specific heat capacity variation in the recuperators (Liu et al.
2019).

Cascaded Cycle Configurations


A big limitation of both the simple and recompression sCO2 cycles is that they
utilize heat from an external heat source over a small temperature interval.
The CO2 is typically heated up to temperatures above 300◦ C in the recupera-
tor, meaning that all available heat from the heat source at lower temperatures
cannot be used in the power cycle. Adding a bottoming sCO2 cycle with a
lower TIT than the topping cycle would allow further utilization of the avail-
20 CHAPTER 2. LITERATURE REVIEW

able heat from the heat source. This type of cascaded cycle configurations
was first proposed by Johnson et al. (2013) to minimize the thermal storage
salt inventory in solar power applications. Manente & Lazzaretto (2014) in-
vestigated the cascaded cycle configurations for sCO2 cycles for electricity
production from biomass. Their research looks at the feasibility of either hav-
ing a simple recuperated cycle or recompression cycle as the topping cycle,
with a simple recuperated cycle as the bottoming cycle. The cycle configura-
tions can be seen in Figures 2.9 and 2.10.

Figure 2.9: Recompression as Topping Cycle.

Figure 2.10: Simple Recuperated as Topping Cycle.

Preheating Cycle
The preheating cycle is essentially the simple recuperated cycle except a frac-
tion of the CO2 flow coming out the the compressor is sent to a preheater with
the remainder of the flow goes through the recuperator. The two flows are then
directly combined and enter the primary heater. The cycle configuration can
be seen in Figure 2.11. This approach has two main benefits. Firstly, it can
ameliorate the pinch point problem in the recuperator by having different mass
flow rates on each side in the heat exchanger. Secondly, having both a primary
CHAPTER 2. LITERATURE REVIEW 21

heater and preheater enables the heat source temperature to be lowered fur-
ther compared to the simple recuperated cycle. This benefit being especially
pertinent in relation to waste heat recovery applications, increasing the energy
recovery efficiency at high thermal power conversion efficiency (Wright et al.
2016).

Figure 2.11: Preheating Cycle

Heat Pumps
Another use of sCO2 systems is heat pumping. A sCO2 cycle can be used
to take low grade heat and pump it to a higher temperature when electricity
prices are low. When electricity prices rise again, the high grade heat can be
transformed into electricity.
In a conventional heat pump, the operating temperature range is limited by
low critical temperatures. In other words, heat cannot be delivered at a tem-
perature higher than the critical temperature. With sCO2 , the operation is not
limited by the low critical temperature. This is due to the operating pressures
being higher than the critical pressure and therefore the operating temperature
is not limited by the critical temperature (Singh & Dasgupta 2017).
Starting from compressor. The CO2 is compressed to a higher pressure
and temperature. Then it is used to heat the storage medium (which is at a
22 CHAPTER 2. LITERATURE REVIEW

lower temperature), so its temperature glides down in a heat exchanger in near


constant pressure. Then it goes into an expander/turbine, where some energy
can be recovered to drive the compressor. The fluid expands and temperature
drops slightly. After that it goes into the recuperators to give off heat for heat
recovery followed by the condenser. When electricity prices are high, the pre-
vious process is reversed. The low pressure sCO2 is compressed to a higher
pressure. Then the fluid receives heat from the the storage. Therefore, its tem-
perature rises significantly. It then expands in a turbine to generate electricity.
Then it goes into a condenser to discard the low grade heat before restarting
the cycle.

2.3.2 Cycle Equipment


There are many components that make up the sCO2 cycle. The main compo-
nents are the turbines, compressors and heat exchangers. The cycle efficiency
is highly affected by the turbine and compressor efficiencies. Therefore, it is
necessary to accurately model these efficiencies. On the other hand, heat ex-
changers make up the recuperators, condenser and the waste heat exchanger.
The effectiveness of the heat exchangers greatly affects the cycle performance.
These components are briefly presented here in regards to the sCO2 cycle.

Turbines
Turbines deal with working fluids at very high temperatures and pressures.
The internal flow inside turbines is a complex phenomenon due to the viscos-
ity and compressibility properties of sCO2 . Therefore more research using ex-
periments or complex numerical computations is needed. Not much research
has gone into centrifugal turbines where sCO2 is the working fluid. Different
manufacturers have went different paths to develop them. Toshiba has been
working on axial turbines (Allam et al. 2017). Axial turbines are mostly used
in large scale applications (Zhang et al. 2015). Axial turbines can withstand
higher temperatures and pressures (Weiland et al. 2019). While NET power is
working on developing a radial turbine, so that it would be simpler and with
fewer stages (El Samad et al. 2020).
There are a few technical limitations for sCO2 turbines. First, the density
of sCO2 is high and the size of the turbine can be relatively small, the heat
exchanger needs to be compact to match. Second, the turbine operates in a
high pressure and high speed environment, which can cause large frictional
losses (Beucher et al. 2010).
CHAPTER 2. LITERATURE REVIEW 23

Highest isentropic efficiency was found to be 87% using a centripetal tur-


bine (Liu et al. 2019). However, it is only suitable for small flows. There are
different sCO2 turbines currently being tested, some of which have an effi-
ciency of up to 86% (Liu et al. 2019).
The choice of turbine type will depend on the size, application but mostly
economics. Axial turbines are generally cheaper (Weiland et al. 2019).

Compressors
Compressors are a vital part of the sCO2 cycle. They are the main source
of energy loss, especially in tCO2 cycles (Vinnemeier et al. 2016). In a sCO2
recompression cycle the recompressor can account for around 40% of the total
work input to the system (McTigue et al. 2019). Compressors are classified
into axial and radial categories. Radial compressors can be either centrifugal
or centripetal compressors. Axial compressors are known for handling large
flows and having high efficiencies. Radial compressors are characterized by
having single-stage pressure ratio, compact structure and low cost (Liu et al.
2019).
Since compressors deal with sCO2 close to the critical point, condensation
should be avoided as two-phase flow is harmful to the equipment. Different
researchers have seen that condensation most likely occurs in the leading edge
of the main impeller blades, the blade tip at the leading edge of main blade, the
trailing edge of both of the main and splitter blades, and the leading edge of
the vaned diffuser (Pecnik & Colonna 2011, Pecnik et al. 2012, Rinaldi et al.
2015, 2014, 2013).
Highest isentropic efficiency was found 84% for a centrifugal compres-
sor. The pressure ratio is generally around 1.8 (Liu et al. 2019). Currently, air
compressors are limited to temperatures of 600◦ C. Whereas, compressors for
CO2 are available only up to 450◦ C (Vinnemeier et al. 2016). Therefore much
research is needed to further develop CO2 compressors to match air compres-
sors.
Turbomachinery for CO2 has higher efficiency and lower costs than those
used for air. This is because CO2 has higher density, which additionally allows
for more compact equipment (Mercangöz et al. 2012). Its low surface tension
allows the turbomachinery to operate near saturation curve as cavitation has a
smaller effect.
24 CHAPTER 2. LITERATURE REVIEW

Heat Exchangers
Heat exchangers in sCO2 cycles deal with fluids at very high temperatures
and pressures. they face major mechanical, thermo-mechanical, and thermal-
hydraulic challenges(Carlson 2014). Recuperators enable high cycle efficien-
cies, and avoids the pinch point in the low-temperature end of the recuperation
process (Dostal et al. 2004). They can be a tubular type with a staggered tube
bundle (Seo et al. 2020), or recent research for this type has been focused
on printed circuit heat exchangers (PCHEs) and cast metal heat exchangers
(CMHEs) (Carlson 2014). PCHEs are the most commonly used for sCO2 cy-
cles. They are manufactured by chemically milling channels roughly 1mm x
0.5 mm into plate swith a size of 600mm x 1500 mm (Musgrove et al. 2014).
This produces compact heat exchangers that can withstand the high pressures
required by the sCO2 cycle.
Another parameter that impacts the cycle efficiency is the condenser power
consumption. As mentioned above, one of the benefits of using an sCO2 cy-
cle is the possibility to have dry cooling. However, dry cooling might require
huge fan power. It is also affected by the cooling air maximum discharge tem-
perature. Rankine cycles discharge waste heat air at around 120◦ C (Saadaoui
2020). For Industrial refrigeration, the cooling air has a maximum temperature
of 45◦ C in the EU (Koelet & Gray 2017).
API standard 661 regulates heat exchanger design for use in refinery (Stan-
dard & Edition 2002). However, it only specifies a difference between inlet and
outlet without setting a maximum discharge temperature. It also states that the
temperature should not exceed 60◦ C due to the technical limitations of the fan.
This is also confirmed by KLM technology group (KLM 2011).
Sandia has been using heat exchangers with temperatures that reached
482◦ C (Shiferaw 2017). The effectiveness of the heat exchangers used de-
pends on the economics as heat exchangers form a significant part of the over-
all CAPEX.

Electric Heaters
Since there are technical limitations on the temperature of the sCO2 reached
by the compressor, electric heaters can be used to further heat sCO2 . This is
particularly useful for heat pumps. Using electric heaters lowers the round trip
efficiency, but reduces the technical challenges of designing high temperature
heat pumps (Vinnemeier et al. 2016). The use of electric heaters depends on
the application and the economics.
CHAPTER 2. LITERATURE REVIEW 25

2.3.3 Emissions Trading System


For the purpose of this thesis, several industry players were contacted. One of
the biggest gas compression industries expressed worries of losing free emis-
sion allowances due to electricity generation. To confirm the findings, the
EU ETS commission was contacted and confirmed the conclusions. There-
fore this section, explores how the sCO2 cycle affects the Emission Trading
Scheme (ETS).
The EU ETS Handbook gives guidance in this matter. It states that "For any
allocation to electricity generators the linear reduction factor (LRF) is applied
to the total allocation. The LRF reduces the total allocation annually by 1.74%
compared to the allocation for 2013. Article 3(u) of the EU ETS Directive
defines an installation as ‘electricity generator’ if “on or after 1 January 2005, it
has produced electricity for sale to third parties, and in which no activity listed
in Annex I (of the EU ETS Directive) is carried out other than the combustion
of fuels”" (EU Commission 2015).
However, to be identified as an electricity generator, these 4 conditions
must apply:

• An electricity generator has to be an installation (see also Article 3e of


the revised EU ETS Directive).

• An electricity generator has to produce electricity.

• An electricity generator has produced or produces electricity for sale at


any time from 1 January 2005.

• An electricity generator must not carry out any activity listed in the An-
nex other than the "combustion of fuels". (EU Commission 2019a)

Therefore, if the gas compressor station produces electricity and uses it to


drive an electric compressor, it does not qualify as an electricity generator.
This means that its emission allowances should not decrease.
Another article of interest is "No free allocation shall be made in respect
of any electricity production, except for cases falling within Article 10c and
electricity produced from waste gases" (EU Court of Justice 2019).
The definition of a waste gas in Art.2(11) of the FAR states that:" ‘waste
gas’ means a gas containing incompletely oxidised carbon in a gaseous state
under standard conditions which is a result of any of the processes listed in
point (10), where ‘standard conditions’ means temperature of 273 K and pres-
sure conditions of 101,325 Pa defining normal cubic meters (Nm3 ) according
26 CHAPTER 2. LITERATURE REVIEW

to Article 3(50) of Commission Regulation (EU) No 601/2012 " (EU Com-


mission 2019b). Therefore the waste heat does not qualify as waste gas.
Many installations produce more than one product. In these cases an in-
stallation can be divided into a number of ‘sub-installations’. "The boundaries
of a sub-installation are determined by the benchmark being applied . Use a
product benchmark, a heat benchmark and a fuel benchmark. The allocation
would need to be calculated separately for each sub-installation." (EU Com-
mission 2015)
Since heat is what is being utilized here. The waste heat can be treated
as a product Benchmark Heat. This means that they will get an extra 224.3
Allowances /GWh of heat consumed. Heat should meet the conditions below
in order to qualify as a heat benchmark sub-installation:

• The heat is used for a purpose (production of products, mechanical en-


ergy, heating, cooling)

• The heat is not used for the production of electricity

• The heat is not consumed within the system boundaries of a product


benchmark

• Heat is consumed within the ETS installation’s boundaries and produced


by an ETS-installation (EU Commission 2019a) (article 3(c)).

The definition of electricity production is not clear whether it is the same


as above or not. However, to gain those extra allowances the turbine can turn
the high pressure into mechanical power that drives a gas compressor right
away, rather than generate electricity.
The EU ETS Handbook states that "installations where the emissions are
so small that the administrative costs per unit of emissions might be dispro-
portionately high are allowed to opt-out from the EU ETS as long as they are
subject to equivalent measures. Installations are considered small emitters if
they emit less than 25 ktCO2e annually and, if they are combustion installa-
tions, have a thermal rated input below 35MW. Hospitals may also opt-out if
they are subject to equivalent measures" (EU Commission 2015).
Sweden doesn’t opt-out installations with low total input or low fossil emis-
sions. Sweden has instead chosen to opt in installations as long as they are
connected to a district heating network where the total input to the network is
greater than 20 MW. Sweden has not chosen to opt any installations out of the
ETS. On the contrary, Sweden applies opt-in, thereby including installations
with a thermal rated input of less than 20 MW. In short Sweden has chosen to
CHAPTER 2. LITERATURE REVIEW 27

apply the threshold of 20 MW, which applies on installation level in the EU,
to each distribution network for municipal heating. Any combustion unit, no
matter how small on installation level, is included in the ETS if it is connected
to a network where the total thermal rated input of all installations connected
exceeds 20 MW (Swedish Energy Agency 2019).
In conclusion, the ETS won’t be affected if the electricity is used on site.
The industry can opt out of deductions if input thermal energy is less than
35MWth, and another policy is in place (Aschenbruck et al. 2004). It also
can be part of the compressor station, not all of it. For example, it can be
installed for one compressor, rather than all the compressors in the plant. An-
other option is that the plant can be divided into two sub-installations. The
first compressor will have no change and is now a separate sub-installation.
The Energy-generator is another one, with different allocations. According to
GD2, they can have a sub-installation of heat, as long as mechanical energy is
used right away instead of electricity. They will receive an extra free allocation
of 224.3 Allowances/GWh of heat consumed (EU Commission 2019a). So us-
ing this alternative actually gives them extra allowances rather than decrease
allowances as was feared.
Chapter 3

Industries and Waste Heat Esti-


mation

3.1 Iron and Steel


The production of steel is an energy intensive industry and consists of many
subprocesses. It involves sintering, coke making, furnace and casting. Each
of these processes involves different alternative methods. A lot of research
has went into heat recovery (Jouhara et al. 2018). Most heat is utilized well in
the industry, except in coke production. Coke production is around 10% of the
total energy demand in the industry. The process involves heating coking coals
to 1100 ◦ C for about 21 hours in the coke oven to drive off volatile compounds.
About 3.3% of the total energy used is electricity, 7.4% is steam and 89.3%
is thermal energy from Coke Oven Gas (COG) combustion (Qin & Chang
2017). To cool down the coke it undergoes a quenching process, the industry
uses either coke wet quenching (CWQ) or coke dry quenching (CDQ). For the
heat to be recovered, it must be CDQ. Of the 62 coke production facilities, only
five have CDQ (Wortler et al. 2013). Consequently, 57 facilities can switch to
CDQ to reduce their carbon emissions and water usage. The coke production
process can be seen in Figure 3.1. There are two high grade heat streams
available for power production.
For CDQ, the ratio of waste gas to coke is 1.23 (Qin & Chang 2017). Using
this ratio, a medium sized plant’s flow can be estimated. The EU produces 37
Million tons of steel per year in 62 facilities (Eurostat 2020). This is equivalent
to coke production of 19 kg/s in each facility resulting in 23.2 kg/s of waste
gas flow. The waste heat in the EU is 15 TWh per year. For each facility, it
can produce approximately 9 MWe using an sCO2 cycle. As for the COG, the

28
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 29

Figure 3.1: Coke Production Process

average coke battery produces 365 m3 /tcoal of COG (Ndlovu et al. 2017, Tran
et al. 2016). For an average size facility this is equivalent to 3.65 kg/s with a
density of 0.5 kg/m3 (Satyendra 2015). It has a high specific heat capacity of
3.3 KJ/kg.K (Zhang 2019). For the EU this is waste heat of 4.6 TWh. For each
facility, it can produce approximately 2.5 MW using an sCO2 cycle. There are
two alternatives for how electricity can be produced. It can be a simple sCO2
cycle as shown in Figure 3.3. Otherwise, it can be a cascaded cycle, where
more waste heat is utilized as seen in Figure 3.2.

Figure 3.2: Combined 2 Cycle for Coke Production


30 CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION

Figure 3.3: Simple sCO2 Cycle for Coke Production

3.2 Gas Compression


There is no official aggregate data for the gas compression industry in the EU.
However, a group of researchers gathered data about the industry (Campana
et al. 2013). The distribution of gas compressor stations in the EU is shown
in Figure 3.4. For the work of this thesis, the group is contacted to see how
they gathered the data. They went through different companies’ databases one
by one to gather their data. Therefore, their data is used as it is. Bianci et al
investigated a an average size compressor station and the amount of waste heat
there. They concluded that the flue gas had a flow rate of 69 kg/s and temper-
ature of 540◦ C (Bianchi et al. 2017). According to Campana et al, the average
useful waste from a station is 84.1MWth . However, the researchers multiply
by a correction factor of 0.65 to account for backup units. Therefore the av-
erage useful waste is 55MWth . This means that the flow rate for a medium
sized station is 126 kg/s. There are approximately 254 stations with different
sizes. The total available waste heat in Europe is 127 TWh. However, only
22.7 TWh are useful. If the gas compressor station has a thermal efficiency of
0.4, then 5.5 GW of electricity can be generated from all the stations.
However, gas compression isn’t a constant load. As can be seen in Figure
3.5, it changes throughout the year. This is also confirmed by other researchers
(Bianchi et al. 2017, ROSSIN 2013). This variation is accounted for in the
results. Similar to the coke production process, this can either be a simple
cycle or a combined cycle.
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 31

Figure 3.4: Distribution of Gas Compressor Stations over Europe


(Campana et al. 2013)
32 CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION

Figure 3.5: Variation of Flow Rate throughout the Year


(Gómez-Aláez et al. 2017)

3.3 Cement
For the Cement industry, a survey was sent out to different factories over Eu-
rope. The survey is attached in Appendix C. Data from 3 main industrial play-
ers was used. Each industrial player represent a different factory size. The
factor of heat per production is almost equal in all of them. However, different
factories use different compositions of fuel. Their numbers are compared and
confirmed using an IFC report (Doğan et al. 2018). The waste heat comes out
at 300 ◦ C with a flow rate of 74 kg/s from the kiln and 40 kg/s from the clinker
cooler. It has a specific heat capacity of 1300 (J/kg K ). The cement manufac-
turing process is seen below in Figure 3.6. The EU produces 125-200 Mtons
of clinker per year (de Vet et al. 2018, Cembureau 2019). This is equivalent
to 120 medium sized factories. They have a cumulative energy of 27 TWh of
waste heat energy.
As seen from the figure above, the waste heat is utilized in the process.
So there are several methods that the sCO2 cycle can add value here. It can
take the waste heat from the process at 285 ◦ C, and use it as the heat source.
However, this will have a low thermal efficiency since the Turbine inlet tem-
perature is relatively low. This can generate 0.5 GW in Europe. Several plants
have been implemented globally, the first was in Germany and then 120 more
power plants were built in China with a capacity of 700 MW (Moya et al.
2010). Another alternative is that before the heat goes into the preheaters, it
passes through a heat exchanger that absorbs some of the high grade waste
heat into the sCO2 cycle and then the rest of the heat goes into preheating.
This schematic can be seen in Figure 3.7. Through this process, the EU can
generate 5 GW of electricity.
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 33

Figure 3.6: Clinker Forming Process

Figure 3.7: Upstream sCO2 Cycle in Cement Industry


34 CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION

Figure 3.8: Cement sCO2 Heat Pump

There are two ways to ensure the heat taken does not adversely affect the
preheating process. First approach is to burn more fuel in the rotary kiln.
Second approach is to replace the 4-stage preheater with a 6-stage preheater
so that the kiln feed absorbs more heat. This would cost A C5-8M (CSI 2017).
Ma et al. (2019) have shown that using the existing preheaters, there is a lot of
room for better thermal exchange and heat utilization.
The last alternative is to use the cycle as a heat pump and engine. When
the electricity prices are cheap, the cycle pumps up the low grade heat and
stores it in a backed bed. Then when the prices are high, the cycle is reversed,
and the heat stored is used to generate electricity using the sCO2 cycle. This
process can be seen in Figure 3.8 and 3.9.

3.4 Aluminum
Aluminum has several different waste streams (Yu et al. 2018). According
to (World Aluminium 2020), the EU produces 7.5 Million tons of Aluminum
every year. The production of Aluminum is a complex process, but it mainly
consists of two main subprocesses; Bayer and Hall-Heroult processes. In the
Bayer process, the Bauxite is crushed and changed into Alumina using heat
mainly that can go up to 1100 ◦ C. The Hall-Heroult process is fundamentally
electrolysis at 940-980 ◦ C in molten cryolite, to produce molten aluminum.
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 35

Figure 3.9: Cement sCO2 Heat Engine

Molten aluminum is then cast or mixed with other elements to form alloys, it
operates 24 hours. Recycling Aluminum is less energy-intensive, this is why
using recycled Aluminum is one of the main goals of 2050. There is also
research into alternative manufacturing methods but nothing will be commer-
cialized before 2030 (CTCN 2012).

3.4.1 Hall Heroult


The Hall Heroult process requires mainly electricity and keeps the heat con-
stant. Smelters consumer 13MWh of electricity per ton of Aluminum in elec-
trolysis (Nowicki & Gosselin 2012). There are two heat streams in this pro-
cess; the exhaust flue gas and the pot surface. The exhaust flue gas has a tem-
perature of 110◦ C but can be slowed down to increase the temperature up to
150◦ C (Yu et al. 2018). For example, a smelter in Iceland of 350,000 t/annum
capacity can have a flow rate of 230 kg/s at that higher temperature (Yu et al.
2018). Zhao has analyzed slowing the pot cooling air to increase its tempera-
ture (Zhao 2015). The main concern is this could melt the frozen electrolyte
and jeopardize pot integrity. Therefore, it is not recommended because of the
subtle thermal balance between the electrolytic bath and the heat loss from
sidewalls. Sørhuus & Wedde (2016) developed a heat exchanger which has a
good trade-off between heat recovery and cost efficient cooling of pot gas.
36 CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION

3.4.2 Bayer Process


In the Bayer process, the input is mostly heat. It requires 4 MWh of heat and
0.15 MWh of electricity (The International Aluminium Institute n.d.). It is
estimated that 30-45% of waste heat in is carried away in the exhaust at 120-
190◦ C (Yu et al. 2018, Paspaliaris et al. 1999). Brough et al estimate that the
exhaust gas carries 0.2 MWh/ton at 160◦ C (Brough & Jouhara n.d.). The flow
rate of the exhaust is slower than the Hall-Heroult, therefore it is lower. The
heat is utilized for preheating. The preheaters are in constant development.
The 5th generation preheaters consist of two stages of preheating (Scarsella
n.d.). Ying Ying et al. studied an Alumina factory in China that had a size
similar to an average European factory (Yingying et al. 2015) and measured
the flow rate (57.4 kg/s) and temperature of the flue gas (160◦ C) . EU produces
6.8 Mtons of Alumina in 5 active facilities. The production capacity of these
facilities is similar to the production capacity Ying et al studied (Fraunhofer
Institute 2009). Below in Table 3.1 is a summary of the waste heat from the
industry. The waste heat is multiplied by the EU production to calculate the
potential in Europe. The highest potential is of low grade waste heat.

3.4.3 Anode Baking


For every ton of Aluminum, 0.4 tons of Carbon electrodes are used (Tajik et al.
2017). The process is shown below in Figure 3.10 along with a graph of tem-
peratures in Figure 3.11. Carbon anode manufacturing consists of aseries of
processes including petroleum coke calcining, crush and grinding of calcined
coke, paste mixing, green anode forming, cooling and anode baking (Qingcai
et al. 2012). The temperature goes up to 1000 ◦ C. There are 18 facilities in the
EU (Fraunhofer Institute 2009) with a production capacity of 2.4 Million tons.
Different sources state that the flue gas comes out at around 300 ◦ C (Becker
n.d., Linhares et al. 2016). This can be used for a downstream sCO2 cycle,
however the mass flow is too low. No upstream cycle can be implemented
there.

3.4.4 Furnace Chimney


Albert analyzed the industry and concluded that the flue gas is very variable
in terms of temperature and flow rate. It changes between 288-1100 ◦ C and
50-14,500 m3 /hr as showm in Figure 3.12. However, the flue gas is used for
preheating, and therefore usually leaves at 300◦ C. It is too transient to use for
an sCO2 cycle.
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 37

Figure 3.10: Schematic of Anode Baking


(Tajik et al. 2017)

Figure 3.11: Temperatures along the Anode Baking Furnace


(Tajik et al. 2017)
38 CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION

Figure 3.12: Variations along the Furnace


(Albert 2012)

Table 3.1: Summary of Aluminum Industry and extrapolating to Europe


Process Heat Potential EU Potential Obstacles
3MWh/t Contaminants and
Hall Heroult Flow 9.8 TWh
100-120◦ C fouling
3MWh/t Maintaining the
Pot Surface Heat Flux ◦ 2.4 TWh
150-200 C heat balance
0.22 MWh/t
Bayer Exhaust Flow 1.5 TWh 3/8 have WHR
110-190◦ C
0.3MWh/t
Anode Baking Sector 0.35 TWh Low thermal output
250-300◦ C
200-1100◦ C
Furnace Chimney 0.65 TWh Transient
0.15 MWh/t
CHAPTER 3. INDUSTRIES AND WASTE HEAT ESTIMATION 39

Figure 3.13: Aluminum Upstream sCO2 Cycle

As mentioned above, Anode baking and furnace heat cannot be used for
an sCO2 cycle. The heat of the Hall-Heroult process should not be meddled
with. Therefore, only the Bayer exhaust flow can be used. The temperature
after preheating is too low to use for an sCO2 cycle. Two ways that the sCO2
cycle can add value is through an upstream integration shown in Figure 3.13,
and heat pumping. The EU can benefit 25 MW from upstream integration.

3.5 Waste Heat Estimation Summary

Table 3.2: Summary of Waste Heat Estimations in the EU


EU Waste Heat (TWh) Power Capacity (GW)
CDQ 0.6
20
CDQ Combined 0.9
GCS 4.125
122
GCS Combined 5.5
Cement Upstream 5
Cement Downstream 27 0.5
Cement Heatpump 0.5
Aluminum Upstream 0.025
1.5
Aluminum Heatpump 0.025
Chapter 4

Performance Model

For the development of the sCO2 technology, modeling and simulation are
essential to support decisions of investment and development. Modeling is
necessary in order to predict the outcomes in terms of energy, annuities, oper-
ation and economics. The functionality of the model will be demonstrated for
the sCO2 recompression cycle shown in Figure 4.1.

Figure 4.1: sCO2 Recompression Cycle

4.1 System Optimization


The techno-economic model aims to optimize the system performance in order
to maximize the Net Present Value (NPV) of the waste heat recovery system.
The optimization of thermal systems involves a large number of decision vari-
ables and constraints. Conventional methods for this optimization apply an
iterative procedure which may lead to solutions at local optimum. Advanced
optimization algorithms, such as evolutionary and swarm intelligence-based

40
CHAPTER 4. PERFORMANCE MODEL 41

algorithms, offer solutions to this problem. They are able to find a solution
closer to the global optimum, with reasonable computational costs. Patel et al.
(2019) performed optimizations for various thermal systems using eleven of
the most popular advanced optimization algorithms. They concluded that for
the Brayton Power Cycle an algorithm called Artificial Bee Colony (ABC) was
best suited for optimization of the system. The techno-economic model in this
thesis will therefore use that optimization algorithm for the sCO2 Brayton Cy-
cle.
The ABC algorithm was proposed by Karaboga (2005) and then further
developed by Karaboga & Basturk (2007). It is nature inspired based on the
foraging behavior of honeybee swarms, mimicking the insect’s food searching
ability. In the algorithm, the artificial bee colony is made up of three groups
of bees, namely, employed bees, onlookers and scouts. The position of a food
source represents a possible solution to the optimization problem in the algo-
rithm, and the amount of nectar in a food source represents the quality (fitness)
of the associated solution. The colony is divided into two groups, the first half
being onlooker bees and the second being employed bees. In the group of
employed bees, each corresponds to a specific food source which memorizes
the position of that food source. The employed bees search their neighbouring
region to seek better food sources. Then, the new food sources are updated
and shared with the onlooker bees. The onlooker bees have exploitation fea-
ture added to them by means of a method called roulette wheel selection, i.e.
each onlooker bee chooses a food source with a probability related to its fitness
value. That selected food source is then evolved to search for a better position,
similar to the search conducted by the employed bees. After a specified num-
ber of failed trials by the employed and onlooker bees to search for a better
food source, that location is considered to have a poor position that will be
abandoned. To replace the abandoned food source a scout bee is sent to ex-
plore the search space in order to find a new randomly generated food source.
The algorithm iteratively sends the three groups of artificial bees to search the
solution space until a termination criterion is met (Patel et al. 2019).
Figure 4.2 demonstrates the logic flow of the techno-economic optimiza-
tion model. It starts by reading inputs about the waste heat flow fluid properties
along with the cooling medium temperature. The inputs are summarized in Ta-
ble 4.1. Then, an initial population of employed bees is generated, i.e. a set
of various design variables is generated. The design variables and their con-
straints are summarized in Table 4.2. The design variable for the temperature
difference between the heat source (THS ) and the Turbine Inlet Temperature
(TIT) is dependent on which heat source is being utilized. Therefore, equa-
42 CHAPTER 4. PERFORMANCE MODEL

Table 4.1: Information needed for the waste heat flow


Input Symbol
Waste heat temperature THS
Waste heat mass flow rate ṁHS
Waste heat fluid specific heat capacity cp,HS
Cooling fluid temperature TCS

Table 4.2: Ranges of design variables for recompression cycle optimization


Design Variable Lower Bound Upper Bound
High pressure level (Ph ), MPa 18.0 25.0
Low pressure level (Pl ), MPa 7.38 12.5
Split ratio (SR) 0.5 1.0
Effectiveness of HTR (ηHT R ) 0.550 0.999
Effectiveness of LTR (ηLT R ) 0.550 0.999
Primary heater approach
∆THS,lb ∆THS,ub
temperature (∆THS ), ◦ C
Main compressor inlet temperature (T1 ), ◦ C 32.0 50.0
Temperature difference between heat source
4 200
outlet and Primary heat CO2 inlet (∆TM H,out ), ◦ C

tions (4.1) and (4.2) were used to calculate the lower and upper bounds of that
design variable respectively. They ensure that the minimum approach temper-
ature within the waste heat exchanger does not go below 4 ◦ C.
(
THS − 700 if THS > 705
∆THS,lb = (4.1)
4.0 if THS ≤ 705
(
THS − 300 if THS > 305
∆THS,ub = (4.2)
THS − 130 if THS ≤ 305
The maximum value of 700◦ C for the turbine inlet temperature is due to
the limitation of the turbine cost scaling model developed by Weiland et al.
(2019).
The subroutine of solution assignments for the various bee groups, shown
in Figure 4.2, represents the evaluation of a food source that an individual bee
has found. The subroutine uses generated design variables for a food source
and calculates the solution, or fitness, at the particular location. This subrou-
tine will be described in detail in the following section.
CHAPTER 4. PERFORMANCE MODEL 43

Figure 4.2: Optimization Model


44 CHAPTER 4. PERFORMANCE MODEL

4.2 Design Point Model


This section describes the general thermodynamic principles for individual
components in the sCO2 power cycle shown in Figure 4.1. Thereafter, the
logic flow of the techno-economic model is demonstrated.

4.2.1 Turbomachinery
The design point performance of the turbine and compressors are modeled
assuming adiabatic operation with a constant isentropic efficiency (ηis ). The
fluid enters the turbomachinery with the specific enthalpy and entropy (hin and
sin ), which are determined using the known temperature and pressure at the
inlet. The isentropic specific enthalpy at the outlet (hout,is ) is then determined
based on the outlet pressure and inlet specific entropy. The isentropic specific
work (wis ) of the turbomachinery can now be calculated using equation (4.3).

wis = hin − hout,is (4.3)


Using the definition of isentropic efficiency, the actual specific work (w)
for a turbine and a compressor can then be calculated using equations (4.4)
and (4.5) respectively.

wturbine = wis · ηis (4.4)

wis
wcompressor = (4.5)
ηis
Finally, the specific enthalpy at the outlet can be calculated using equation
(4.6).
hout = hin − w (4.6)
With the calculated outlet specific enthalpy and the known outlet pressure,
the thermodynamic state at the outlet of the turbomachinery is fully defined.

4.2.2 Heat Exchangers (HEXs)


As can be seen in Figure 4.1, the recompression cycle consists of three dif-
ferent types of heat exchangers (two recuperators, primary heater and cooler).
In order to model the thermodynamic cycle performance and the economics
of the system, the conductance (UA) along with the thermodynamic states of
inlets and out lets for each heat exchanger need to be determined. One of the
CHAPTER 4. PERFORMANCE MODEL 45

challenges and interesting properties of sCO2 is the change of its fluid prop-
erties around the critical point. Therefore, normal heat transfer correlations
cannot be used for the modeling of an sCO2 system (Cabeza et al. 2017).
The model accounts for pressure losses across the heat exchangers. There
are different methods to account for pressure losses. The first method is to
go into the details of the heat exchanger piping geometry and computationally
simulate the fluid flow. However, this method is very computationally expen-
sive and therefore pressure losses can be based on the literature. Wright et al.
(2016) have estimated pressure losses of 1% before the turbine and 2 % after
the turbine. Dyreby (2014) assumed that the pressure drop across each heat
exchanger is 1%. While, Le Moullec (2013) modelled a coal fired plant using
sCO2 . They calculated a pressure drop of 1.2% before the turbine and 2.5%
after. Since that system is more complex with more heat exchangers, Dyreby’s
estimate of 1% across the entire heat exchanger relative to the inlet pressure
on either side is used. This section will explain each heat exchanger module.

Recuperator
The recuperators are modeled assuming a counter-flow configuration. Three
main working principles are utilized to calculate all four thermodynamic states
at both sides of the recuperators. Firstly, the law of energy conservation across
the HEX is used to calculate the final state when three are fully defined. Sec-
ondly, the definition of heat exchanger effectiveness, and lastly, the log mean
temperature difference (LMTD) approach to determine the conductance of the
HEX.
The fluid enters the hot and cold side of the recuperator with specific en-
thalpy (hh,in , hc,in ) and exits the hot side with specific enthalpy (hh,out ), all
of which are determined from known temperature and pressures. Using the
law of energy conservation the cold side outlet specific enthalpy (hc,out ) can
be determined as shown by equation (4.7)

ṁh · (hh,in − hh,out )


hc,out = hc,in + (4.7)
ṁc
For a recuperator that has the same fluid on both hot and cold sides this
equation can be simplified. For the HTR the mass flow rate is equal on each
side and therefore cancel out. Due to the split flow in the recompression cycle
the mass flow rates are not equal but their ratio is known and is equal to the
split ratio (SR). Equation 4.7 is simplified to:
(hh,in − hh,out )
hc,out = hc,in + (4.8)
SR
46 CHAPTER 4. PERFORMANCE MODEL

From the determined specific enthalpy and known pressure losses across
the HEX all thermodynamic states are fully defined.
Equation (4.9) shows the definition of heat exchanger effectiveness.

Actual heat transfer q


 = = (4.9)
Maximum possible heat transfer qmax
The actual heat transfer may be determined from either the energy lost by the
hot fluid or the energy gained by the cold fluid. Equation (4.10)

q = ṁh ch · (Th,in − Th,out ) = −ṁc cc · (Tc,in − Tc,out ) (4.10)

To calculate the maximum possible heat transfer, the maximum tempera-


ture difference present in the HEX is used, which is the difference between the
inlet temperatures for the hot and cold fluids. Furthermore, the fluid that might
undergo the maximum temperature difference is the one with the minimum
value of ṁc. The maximum possible heat transfer can then be determined
using equation (4.11).

q = (ṁc)min · (Th,in − Tc,in ) (4.11)


By selecting the appropriate fluid side to calculate the actual heat transfer,
the effectiveness can be determined with only knowing the temperatures across
the HEX. Equation (4.12) shows a general way of expressing the effectiveness
this way.

∆T (Minimum fluid)
 = (4.12)
Maximum temperature difference in heat exchanger
The minimum fluid is always the one undergoing the larger temperature
change in the heat exchanger. The maximum temperature difference in the
heat exchanger is always the temperature difference of the hot and cold fluid
inlets (Holman 2018). This enables the effectiveness to be calculated without
knowing the mass flow rates of the working fluid. Equally important, this
allows for the thermodynamic properties at every point along the cycle to be
determined.
The heat transfer rate across a heat exchanger can be expressed using the
LMTD method according to equation (4.13).

Q̇ = (U A)∆TLM T D (4.13)
CHAPTER 4. PERFORMANCE MODEL 47

where ∆TLM T D for a counter-flow configuration can be expressed as,

(Th,out − Tc,in ) − (Th,in − Tc,out )


∆TLM T D =  
T −Tc,in
ln Th,out
h,in −Tc,out

However, this method involves two important assumptions. First one is


that the specific heats of the fluids do not vary with temperature, and the sec-
ond that the convection heat transfer coefficients are constant throughout the
heat exchanger. In the case of sCO2 cycles, both these assumptions are likely
to be broken due to the large variations in thermodynamic properties of CO2
under supercritical conditions. Therefore, to accurately capture the effects of
changing fluid properties, each heat exchanger is discretized into sub-section
connected in series (Nellis & Klein 2012). Using this approach, the total con-
ductance to a given recuperator is determined using the known inlet and outlet
conditions. The total heat transfer rate (Q̇) through the whole heat exchanger
is calculated and evenly distributed amongst the discretized sub-sections. As-
suming that the pressure losses across the HEX are linear the inlet and outlet
states for each sub-section can be fully defined. Using equation (4.13) the
conductance of each sub-section can be determined:

Q̇i
(U A)i = (4.14)
∆TLM T D,i
The total conductance for the recuperator is then the sum of all sub-section
conductance values.
X
(U A) = (U A)i (4.15)

Primary Heater
Using equations (4.3) - (4.12), the thermodynamic properties of every point
in the sCO2 cycle can be determined. Using the design variable of the temper-
ature difference between heat source outlet and primary heater CO2 inlet, the
waste heat flow outlet temperature can be determined. From the known fluid
properties of the waste heat, the total heat transfer rate of the heat exchanger
can be calculated. An energy balance across the heat exchanger can then be
used to calculate the mass flow rate of CO2 . Following the same approach of
the recuperators, the total conductance of the primary heat exchanger can then
be determined.
48 CHAPTER 4. PERFORMANCE MODEL

Cooler
An air cooling system for Brayton cycles conventionally has cross-flow con-
figuration. Since a cross-flow heat exchanger has two fluid flows in perpen-
dicular directions, fluids in parallel channels will observe opposite side fluids
of different temperature. This means that numerical modeling of this system
needs to be two-dimensional, resulting in huge increase in computation cost
(Moisseytsev et al. n.d.). Therefore, to improve computation time, a simplifi-
cation is made by assuming an approach temperature of 15◦ C and calculating
the heat exchanger conductance using the known temperatures on the CO2
side.

4.2.3 Cycle Model


The operating assumptions for the sCO2 waste heat recovery cycles are shown
in Table 4.3.

Table 4.3: Operating assumptions for the sCO2 Waste Heat Recovery cycles.
sCO2 Assumption Symbol Value
Compressor isentropic efficiency ηcomp 89%
Turbine isentropic efficiency ηturbine 93%
Generator efficiency ηgen 98%
Cooling fluid inlet temperature TCS 25 ◦ C
Compressor COP COP 1.5

A flow diagram of the model iteration logic can be seen in Figure 4.3. The
approach follows a similar approach as Dyreby (2014) aside from the different
design variable inputs. These alterations were made for optimizing waste heat
recovery applications specifically.
CHAPTER 4. PERFORMANCE MODEL 49

Figure 4.3: Iterative process logic flow for the cycle model.
50 CHAPTER 4. PERFORMANCE MODEL

Figure 4.3: Continued.

During this iteration process it is possible for a temperature cross-over to


occur, a condition where the hot side temperature drops below the cold side
temperature in a particular region on a particular iteration. This occurrence
implies that the results for this iteration are non-physical due to a violation of
the Second Law of Thermodynamics. This situation is handled by checking
whether all conductance values are real numbers. If not, the fitness function
returns a value of negative infinity so that the optimization algorithm ignores
that solution.
Like the model developed by Dyreby (2014), a combination of the secant
and bisection methods are used for both iteration loops to adjust T7 and T8 .
Even though the secant method has a higher rate of convergence, it can poten-
tially predict new values outside of valid bounds, leading to divergence of the
method. By reverting back to the bisection method when the secant method
fails this can be avoided (Canale & Chapra 2009).
CHAPTER 4. PERFORMANCE MODEL 51

4.2.4 Fluid Properties


The open source C++ library named CoolProp was used to obtain fluid prop-
erties of CO2 (Bell et al. 2014). The CoolProp consists of pure and pseudo-
pure fluid Equations of State (EOS) and transport properties for various com-
pounds. The fluid properties of CO2 calculated by CoolProp are based on
Helmholtz energy formulations (Span & Wagner 1996). While using the EOS
to calculate the fluid properties is extremely accurate, it is computationally
expensive. In order to increase the speed of the optimization model, the tabu-
lar interpolation feature available in CoolProp was used. The model uses the
Bicubic interpolation method.

4.2.5 Economic Model


The economic assumptions used to calculate the cycle’s KPIs for the optimiza-
tion algorithm are shown in Table 4.4.

Table 4.4: Economic Assumptions Used in Calculating KPIs


Economic Assumptions Symbol Value
Nominal Discount Rate i 6%
Lifetime N 25 years
Annual Operating hours Nh 7446 hours
Electricity Price elprice 0.065 A
C/kWhe

Industrial electricity prices vary in different countries from A


C51.8-122.7
per MWh as can be seen in Table 4.5 (BMWi 2017, Industrial electricity prices
across Europe 2018 | Statista n.d.).
Factors of industrial activity of different EU countries are used to calcu-
late a realistic EU nominal price (Industrial production statistics - Statistics
Explained n.d.). Based on that, it was discovered that the nominal industrial
electricity price is 8.5 cA
C/kWh.
The assumed electricity price was based on this analysis along with the
replies from surveys sent to the cement industry as seen in Appendix C.
Weiland et al. (2019) recently developed a cost scaling model for the com-
ponents in a sCO2 cycle. The model is based on a total of 129 vendor quotes,
and spans cycle size ranges of 5-750 MWe . It uses an appropriate scaling pa-
rameter (SP) for different components and includes a temperature correction
facor (fT ) for certain components to account for the increase in cost at higher
52 CHAPTER 4. PERFORMANCE MODEL

Table 4.5: Industrial Electricity Prices in EU Countries


(BMWi 2017, Industrial production statistics - Statistics Explained n.d.)
Country Elec Price c/kWh Share of EU’s Industrial Activity
Austria 7.65 2.75%
Czechia 6.41 2.75%
Denmark 7 2.75%
France 6.35 12%
Germany 8.7 28%
Greece 7.67 2.75%
Hungary 7.07 2.75%
Italy 9.6 16%
Norway 6.41 2.75%
Poland 7.36 5%
Slovakia 10.06 2.75%
Spain 8.79 8%
Sweden 5.18 2.75%
United Kingdom 12.27 9%

temperatures. The cost scaling can be expressed in a general way using equa-
tion (4.16).

CE = a · SP b · fT (4.16)
where,
(
1 if Tmax < Tbp
fT = 2
1 + c · (Tmax - Tbp ) + d·(Tmax - Tbp ) if Tmax ≥Tbp

and CE is the equipment cost for an individual cycle component, a is the ref-
erence cost, and b is the cost exponent in order to take into account economy
of scale. All cost functions along with the reference costs and exponents are
found in Appendix A.
The direct equipment capital cost (CDE ) is calculated according to Bailie
et al. (2018), and can be expressed on a general form as equation (4.17).

CDE = CE · (1 + αM )(1 + αL ) (4.17)


where αM and αL are the multiplication factors for installation cost of ma-
terials and labor respectively. These factors are shown in Appendix A. System
CHAPTER 4. PERFORMANCE MODEL 53

piping costs can vary anywhere from (5-20)% of total power block capital
costs, depending on the cycle operating conditions (White et al. 2017). For
the analysis performed in this thesis a value of 10% was used. Additionally,
direct capital accounting for improvements to site, instrumentation and con-
trols, and other Miscellaneous balance of plant systems are also added (White
et al. 2018).
Indirect costs include Engineering, Procurement and Construction cost
(EPC) along with contingencies. The EPC costs are assumed to be 9% of
the total direct capital cost. The contingencies depend on the status of the
technology being considered and are assumed to be 30% (Gerdes et al. 2011).
Operation and maintenance (O&M) costs include taxes, maintenance ma-
terial costs, and labor costs that account for operating, maintenance, admin-
istrative and support labor (Weiland et al. 2017). A correction factor of 25%
was included to exclude the costs associated with the coal gasification section
of the power plant modeled by Weiland et al. (2017). The indirect costs and
O&M costs are explained in further detail in Appendix A.

4.3 Heat Pump Model


For the heat pump modeling, the same process and equations are used but in
reverse. The heat pump requires a storage system to store the heat. Using a
packed bed thermal storage system overcomes the temperature limits of molten
salts, discussed in section 4.6, which results in achieving a higher round-trip
efficiency.
There are different ways to operate the heat pump. There are more complex
solutions which yield higher round trip efficiencies but cost more (Morandin
et al. 2013). For example, one can use hot and cold storage. Another solution
would be having double tanks or recuperated storage as shown in Figure 4.4.
54 CHAPTER 4. PERFORMANCE MODEL

Figure 4.4: Recuperated Storage


For feasibility, a simple heat pump and storage model as shown in Figure
3.8 is used. So that the aforementioned model is just reversed and used.

4.4 Packed Bed thermal Storage Model


The thermal storage has been chosen to be a packed bed due to its high en-
ergy density, good heat transfer between the air and the rocks, stability and
low cost (Anderson et al. 2015). Based on Schumann (1929)’s model, a one
dimensional, two phases, transient model considering conduction, convection
and radiation heat transfer was summarized by Trevisan et al. (2019). How-
ever, it assumes that no radial heat conduction, nor conduction in the fluid or
solids happens (Anderson et al. 2015). It is assumed that air acts as the heating
fluid, as the sCO2 has a very high pressure that would require costly storage
vessels. The user inputs the storage size needed. Then the code calculates
the heat exchange and the separate components needed for that specified size
along with their costs. The equation can be summarized as follows:
CHAPTER 4. PERFORMANCE MODEL 55

∂TF G ∂TF kFef f ∂ 2 TF has Uw DT ES π


+ = + (TS − TF )+ (T∞ − TF )
∂t ρF ε ∂x ρF cpF ε ∂x2 ρF cpF ε ρF cpF AT ES ε
(4.18)

∂TS kSef f ∂ 2 TF hv as
= 2
+ (TF − TS ) (4.19)
∂t (1 − ε)ρs cpS ∂x ρs cpS (1 − ε)

6(1 − ε)
as = (4.20)
dp
nins
1 1 dinside DT ES X 1 dj+1 1 doutside
= + ln + (4.21)
Uw ai DT ES 2 j=1 kj dj ao dn+1

700
hv = G0.76 dp0.24 (4.22)
6(1 − ε)
Where hv is the convective heat transfer rate between the HTF and the filler
material (Coutier & Farber 1982).
The density and specific heat of the rocks (igneous acid rocks) is retrieved
from (Tiskatinee et al. 2017, Tiskatine et al. 2017, Becattini et al. 2017), while
the rock conductivity is retrieved from (Haenel et al. 2012). These parameters
are summarized in Table 4.6.

Table 4.6: Thermodynamic Properties of Packed Bed Material (Tiskatinee


et al. 2017, Tiskatine et al. 2017, Becattini et al. 2017, Haenel et al. 2012).
Storage Parameters Symbol Value Unit
Density ρs 2850 kg/m3
Specific Heat Cps 172.5 J/(kg K)

Thermal Conductivity ks (807/(350+T[ C]))+0.64 W/(m.K)
Emissivity  0.85 -
Particle Mean Diameter dp 0.035 m

4.5 Key Performance Indicators


To track the performance and feasibility of implementing sCO2 for waste heat
recovery, the following Key Performance Indicators (KPIs) are used:
56 CHAPTER 4. PERFORMANCE MODEL

Table 4.7: Values of the main parameters used for validation of the packed bed
model by Trevisan et al. (2019).
Parameter Symbol Value Unit

TES Inlet Temperature Tin 500 C

Ambient Temperature T∞ 20 C
Specific HTF Mass Flow Rate G 0.225 kg/(m2.s)
Packed Bed Height H 1.2 m
Packed Bed Diameter D 0.148 m
Void Fraction  0.4 -
Particle Hydraulic Diameter dp 0.02 m
Solid Density ρs 2680 kg/m3
Solid Specific Heat cps 1068 J/(kg.K)
Solid Thermal Conductivity ks 2.5 W/(m.K)
External Heat Transfer Coefficient Uw 0.678 W/(m2.K)

• Thermodynamic:

– Thermal Efficiency
– Exergy Efficiencies
– Waste Heat Utilization fraction

• Economic:

– Net Present Value


– Discounted Payback Period
– Demand Self-Sufficiency

• Environmental:

– Mitigated CO2 Emissions

4.5.1 Thermodynamic KPIs


Thermal Efficiency
Thermal efficiency for a power cycle is defined the the ratio between electrical
power produced by the cycle (Ṗe ) and thermal power extracted from the heat
source (Q̇th ). This can be expressed on a general form according to equation
(4.23).
CHAPTER 4. PERFORMANCE MODEL 57

Ṗe
ηth = (4.23)
Q̇th

Waste Heat Utilization Fraction (WHU)


The waste heat utilization fraction is defined as the ratio of waste heat trans-
ferred to the sCO2 cycle (Q̇W HEX ) relative to what is currently discarded from
the industrial site (Q̇exhaust ). For a fair comparison and due to technical limi-
tations, the same reference state is used for all industries. The reference state
used in this work is the Standard Ambient Temperature and Pressure (SATP)
which is at 25 ◦ C and 101.325 kPa, i.e. heat discarded into the environment
at 25◦ C and 101.325 kPa is heat that has been fully utilized. This fraction can
be expressed with equation (4.24)

Q̇W HEX
WHU = (4.24)
Q̇exhaust
Where Q̇W HEX is the heat transfer rate in the waste heat recovery HEXs and
Q̇exhaust is the available heat in the waste heat exhaust gas.

Exergy Efficiency
The exergy efficiency is defined as the ratio between electrical power produced
by the sCO2 power cycle and available heat in the waste heat exhaust.

Pe
ηex = (4.25)
Qexhaust

4.5.2 Economic KPIs


Net Present Value
Net Present Value is a measure of how profitable an investment is. It is defined
as “the difference between the present value of cash inflows and the present
value of cash outflows (Kenton 2020). A positive NPV shows that the revenues
generated by the project are more than the anticipated costs.

Cashf low
NPV = (4.26)
(1 + i)n
where i is the discount rate adjusted for inflation, n is the number of periods
in the future and Cashflow is the revenue for a given time period. (ECEEE
58 CHAPTER 4. PERFORMANCE MODEL

Discount rates n.d., European Union inflation rate | Statista n.d., European
Union Central bank discount rate - Economy n.d.)

Discounted Payback Period


The discounted payback period shows when the project will reach break even.
It is “the time in which the initial cash outflow of an investment is expected to
be recovered from the cash inflows generated by the investment.” It is based
on the investment’s discounted cash in- and outflows (Kenton 2019).

Investment Cost
DPB = (4.27)
Discounted Annual Cash F lows

Demand Self-Sufficiency
The demand self sufficiency shows how much of the electricity demand needed
by the plant is satisfied by the electricity produced by the sCO2 cycle.

Electricity P roduced by the Cycle


DSS = (4.28)
Electricity N eeded by the P lant

4.5.3 Environmental KPI


CO2 Mitigated
It is a factor that shows how much CO2 emissions has been mitigated. All the
electricity generated by the investment is less electricity generated by power
producers. Therefore, by using the emission factor (EF) of electricity produc-
tion in the EU, CO2 mitigation can be expressed as:

CO2,M = Eprod · EFCO2 (4.29)


Where CO2,M is the amount of CO2 mitigated, Eprod is the annual electric-
ity production and EFCO2 is the emission factor which is the average amount
of CO2 produced per kWh in the EU (0.27 kg/kWh) (IEA 2019).
Chapter 5

Results and Discussion

5.1 Different Industries


Different industries were analyzed in this thesis. The inputs into the model are
mentioned in Chapter 3. In Figure 5.1 one can see the NPV, the waste heat
utilization factor and exergy efficiency of the optimum cycle for each industry.
Figure 5.2 shows the CO2 mitigated and DPB periods per industry. While
Figure 5.3 shows the share of electricity produced to the electricity consumed
for each industry.
The cement and aluminum industry currently utilize some of the waste heat
using preheaters. Currently, a cement facility with 4-stage preheater utilizes
60% of the available waste heat. The addition of a downstream or upstream
sCO2 cycle would result in 84% or 79% waste heat utilization, respectively.
Therefore, as can be seen in Figure 5.1, the downstream and upstream sCO2
cycles utilize 24% and 19% of the waste heat, respectively.
A conventional aluminum production facility utilizes 86% of the available
waste heat with preheaters and adding an upstream sCO2 cycle would result in
95% total waste heat utilization, meaning the power cycle utilizes 9% of the
waste heat, as can be seen in Figure 5.1.

59
60 CHAPTER 5. RESULTS AND DISCUSSION

Figure 5.1: NPVs of the Different Industries

Figure 5.2: DPBs of the Different Industries


CHAPTER 5. RESULTS AND DISCUSSION 61

Figure 5.3: Electric Demand Sufficiency of Different Industries

5.1.1 Heat Pumps


As can be seen above, the Heat Pump configuration yields a negative NPV.
This section studies the difference needed between the buying and selling elec-
tricity prices so that the configuration is positive. As can be seen from Figure
5.4, the difference has to be AC165/MWh between the buying and selling elec-
tricity prices for the Cement heat pump to be barely positive. On average, the
difference is AC40/MWh, which yields an NPV of A C-41M.
62 CHAPTER 5. RESULTS AND DISCUSSION

Figure 5.4: The Effect of Difference in Prices on a Cement Heatpump NPV

Figure 5.5, shows the same for the Aluminum Industry. The difference in
electricity prices has to be at least A
C185/MWh, for the NPV to be positive.

Figure 5.5: The Effect of Difference in Prices on an Aluminum Heatpump


NPV
CHAPTER 5. RESULTS AND DISCUSSION 63

5.2 Different Cycles


To compare between different cycle configurations, the Iron and Steel industry
is used. The different cycle efficiencies and NPVs are shown here. It can be
seen in Figure 5.6 that the preheating cycle has the highest NPV. In addition,
even though the recompression cycle has higher thermal efficiency than the
Brayton cycle, yet it has a relatively lower NPV.

Figure 5.6: NPVs of the Different Cycle Configurations

5.3 Effect of Split Ratio on NPV


Recompression cycle has been gaining a lot of attention in the past years due
to its higher efficiency. Section 5.2 shows that the Brayton Cycle has a higher
NPV than the Recompression cycle. This section investigates the effect of Split
Ratio on the NPV of a Recompression cycle for the Iron and Steel Industry.
As can be seen in Figure 5.7, the NPV decreases as the split ratio decreases.
The thermal efficiency on the other hand increases as the split ratio decreases
until it reaches a maximum. This means that it is better to have a Brayton
cycle than a recompression cycle. As mentioned in section 2.3, recompression
cycles give higher thermal efficiencies. However in waste heat recovery, they
64 CHAPTER 5. RESULTS AND DISCUSSION

yield lower NPVs and exergy efficiencies due to waste heat source utilization
being limited.

Figure 5.7: NPVs of Different Split Ratio in Recompression Cycle

5.4 Effect of Turbine Inlet Temperature


One of the most influential factors to the sCO2 cycle is the TIT. As can be seen
from Figure 5.8, as the TIT increases the thermal efficiency increases linearly.
However, the NPV does not follow the same pattern. The NPV rises with TIT
to 635◦ C, then it decreases again. This is because at higher temperatures, the
costs of the components increase. Therefore the NPV rises to an optimum
temperature and then decreases.
CHAPTER 5. RESULTS AND DISCUSSION 65

Figure 5.8: Effect of TIT on NPV and Efficiency

5.5 Effect of Recuperator Effectiveness


In a similar manner, the recuperator effectiveness affects the sCO2 cycle. Fig-
ure 5.9 shows that the NPV also increases up to a certain point, then it starts
decreasing. This is because at higher effectiveness, the heat exchanger be-
comes too big and expensive. Therefore, it has an adverse effect on the NPV
as the heat utilized is not worth the extra costs.
66 CHAPTER 5. RESULTS AND DISCUSSION

Figure 5.9: Effect of Recuperator Effectiveness on NPV and Efficiency

5.6 System Costs


The powerblock has the highest contribution to the CAPEX as can be seen in
Figure 5.10. Since this is a relatively new technology, the contingencies cost
is also high at 23% of the CAPEX cost.

Figure 5.10: CAPEX Share Among Main Plant Components


CHAPTER 5. RESULTS AND DISCUSSION 67

Figure 5.11 shows the relative cost of the different powerblock components
of the preheating cycle. Results show that the waste heat exchanger for the
CDQ has the highest cost followed by the recuperator. The cost distribution
can be seen in Figure 5.11

Figure 5.11: Share of Different Components in Powerblock Cost

5.7 Sensitivity Analysis


A sensitivity analysis is carried out to determine which factors affect the NPV
mostly. The error ranges of (Weiland et al. 2019) are chosen as the sensitivity
ranges. Results show that the components’ price do not affect the NPV greatly.
This proves that the error possibilities in the cost estimation model will not
greatly impact the NPV. Among the different components, the compressor’s
cost has the greatest impact. Figure 5.12 shows the impact of the change in
price of all the components.
68 CHAPTER 5. RESULTS AND DISCUSSION

Figure 5.12: Sensitivity Analysis of Change in Power Block Price


For the other factors, the electricity price has the highest impact on NPV
as can be seen in Figure 5.13. The full sensitivity analysis can be seen in
Appendix B.

Figure 5.13: Sensitivity Analysis of Change in Electricity Price

5.8 Model Verification


This section aims at verifying the developed model. The verification is carried
out by, first comparing the thermal efficiency calculated to the literature. The
CHAPTER 5. RESULTS AND DISCUSSION 69

Table 5.1: Model verification for cycle thermal efficiency using data reported
by Manente & Lazzaretto (2014).
Design variables and
Manente & Lazzaretto (2014) Present work
performance parameters
Maximum pressure 20 MPa 20 MPa

Compressor inlet temperature 32 C 32 ◦ C
Minimum pressure 7.63 MPa 7.63 MPa

Turbine inlet temperature 550 C 550 ◦ C
Split ratio 0.625 0.625
Compressor efficiency 90% 90%
Turbine efficiency 90% 90%
Mechanical shaft efficiency 98% 98%
Calculated thermal efficiency 44.18% 44.07%
Relative difference - 0.25%

discretization in heat exchanger model was also verified by analyzing the con-
vergence for the calculated conductance with varying number of sub-section
in the each heat exchanger.

5.8.1 Cycle Model Verification


Table 5.1 shows the comparison between the model developed in this work
and results from Manente & Lazzaretto (2014). The small difference in ther-
mal efficiency between the models can be explained by the use of different
thermodynamic properties databases.

5.8.2 Heat Exchanger Verification


The heat exchanger convergence was analyzed using the optimization design
variables and constraints for the CDQ waste heat recovery system. The con-
ductance for every heat exchanger was calculated with varying number of sub-
section from 2-100, and the error relative to having 1000 sub-section was de-
termined. This was then repeated using a set of 10,000 randomly generated
optimization variables. The longest convergence case, along with few random
ones, can be seen in Figures 5.14, 5.15 and 5.16. Using a convergence toler-
ance of 1 · 10−3 had the absolute worst case scenario of convergence at 37 heat
exchanger sub-sections. Therefore, that value was used for the discretization
of the heat exchangers. From this analysis the importance of splitting up the
70 CHAPTER 5. RESULTS AND DISCUSSION

heat exchangers into sub-sections becomes evident. Calculating the conduc-


tance of the heat exchangers from end-point temperatures can result in errors
of up to 68%.

Figure 5.14: The Influence of Number of Subsections on Error of Main Heater


Calculations

Figure 5.15: The Influence of Number of Subsections on Error of Recuperator


Calculations
CHAPTER 5. RESULTS AND DISCUSSION 71

Figure 5.16: The Influence of Number of Subsections on Error of Preheater


Calculations
Chapter 6

Conclusion

The power output thermal efficiency, exergy efficiency, and cost can be used
for assessing the performance of the sCO2 cycle but the relationship between
these parameters was studied in this work. The optimal trade-off between the
economics and thermodynamic performance of an sCO2 cycle was estimated
in this work using a techno-economic optimization structure of the different
cycle parameters. This approach is better than others as it optimizes for the best
NPV cycle and it makes the cycle attainable. To be specific, this paper focused
on waste heat from four industries in the EU. The waste heat in the EU from
these four industries was estimated. Then six different cycle configurations
were compared against each other. The pressure levels of the cycle along with
the recuperators’ effectiveness were optimized for the different industries and
configurations to yield the highest NPV. With the constraints assumed in this
work, the highest NPV for a medium Iron and Steel plant is A C34.6M. This
system had a thermal efficiency of 44% and a discounted payback period of 6
years. It was deduced that recompression cycles are not the highest economic
values for waste heat recovery. It was discovered that the preheating cycle was
the most fit among the cycles analyzed. A diagram was derived to compare the
different cycle configurations, and another to compare between the different
industries. The Heat pump configuration is not financially attractive with the
current electricity prices. It needs a difference of AC165/MWh between the
buying and selling prices.
The turbomachinery needed to operate this cycle is a concern. CO2 has
high density at supercritical conditions, which makes the machinery required
compact. However, this also means higher stresses on the blades, this might
prohibit the use of single shaft machines with few stages.
This study identifies the best cycle configurations and parameters to be

72
CHAPTER 6. CONCLUSION 73

used. However, it did not include the assessment of different working flu-
ids. Different working fluids had very different critical points, which meant
completely different setups. It could have been possible to just operate them
in the supercritical region. However, this would mean the benefits of operat-
ing around the supercritical point would be lost. Some technical aspects of
the components were quickly discussed, however they require more investiga-
tions. More analysis is needed for (a) the cycle operation with variable waste
heat (b) operation at off-design point, (c) the investigation of start-ups, shut-
downs and load changes and (d) dual operation of compressors as turbines and
vice versa. Further research is needed on supercritical cycle systems at large
scale to properly evaluate their commercialization.
Bibliography

Albert, D. (2012), ‘Design of heat recovery system in an aluminium cast


house’, Master of Science Thesis in Engineering and ICT, Norwegian Uni-
versity of Science and Technology .

Allam, R., Martin, S., Forrest, B., Fetvedt, J., Lu, X., Freed, D., Brown Jr,
G. W., Sasaki, T., Itoh, M. & Manning, J. (2017), ‘Demonstration of the
allam cycle: an update on the development status of a high efficiency su-
percritical carbon dioxide power process employing full carbon capture’,
Energy Procedia 114, 5948–5966.

Anderson, R., Bates, L., Johnson, E. & Morris, J. F. (2015), ‘Packed bed ther-
mal energy storage: A simplified experimentally validated model’, Journal
of Energy Storage 4, 14–23.

Angelino, G. (1968), ‘Carbon dioxide condensation cycles for power produc-


tion’.

Angelino, G. (1969), Real gas effects in carbon dioxide cycles, Vol. 79832,
American Society of Mechanical Engineers.

Aschenbruck, E., Beukenberg, M., Blaswich, M. & Bokelmann, H. (2004),


The upgraded power turbine for the industrial gas turbine thm 1304 devel-
opment and first operational experience, in ‘Turbo Expo: Power for Land,
Sea, and Air’, Vol. 41693, pp. 645–651.

Bailie, R. C., Whiting, W. B., Shaeiwitz, J. A., Turton, R. & Bhattacharyya, D.


(2018), Analysis, Synthesis, and Design of Chemical Processes, Pearson.
URL: https:// www.oreilly.com/ library/ view / analysis- synthesis- and/
9780134177502

Baldwin, S., Bindewald, G., Brown, A., Chen, C., Cheung, K., Clark, C.
& Cresko, J. (2015), Quadrennial technology review 2015: Technology

74
BIBLIOGRAPHY 75

assessments–advancing clean electric power technologies, Technical report,


EERE Publication and Product Library.

Becattini, V., Motmans, T., Zappone, A., Madonna, C., Haselbacher, A. & Ste-
infeld, A. (2017), ‘Experimental investigation of the thermal and mechani-
cal stability of rocks for high-temperature thermal-energy storage’, Applied
Energy 203, 373–389.

Becker, F. F. (n.d.), ‘Ring pit furnaces for baking of high quality anodes’, An
Overview. Riedhammer, in .

Bell, I. H., Wronski, J., Quoilin, S. & Lemort, V. (2014), ‘Pure and pseudo-
pure fluid thermophysical property evaluation and the open-source thermo-
physical property library coolprop’, Industrial & Engineering Chemistry
Research 53(6), 2498–2508.
URL: http:// pubs.acs.org/ doi/ abs/ 10.1021/ ie4033999

Beucher, Y., Ksayer, E. V. & Clodic, D. (2010), ‘Characterization of friction


loss in pelton turbine’.

Bianchi, M., Branchini, L., De Pascale, A., Melino, F., Orlandini, V., Peretto,
A., Archetti, D., Campana, F., Ferrari, T. & Rossetti, N. (2017), ‘Techno-
economic analysis of orc in gas compression stations taking into account
actual operating conditions’, Energy Procedia 129, 543–550.

BMWi (2017), ‘Zahlen und fakten energiedaten-nationale und internationale


entwicklung’.

Brough, D. & Jouhara, H. (n.d.), ‘The aluminium industry: A review on state-


of-the-art technologies, environmental impacts and possibilities for waste
heat recovery’, International Journal of Thermofluids .

Cabeza, L. F., de Gracia, A., Fernández, A. I. & Farid, M. M. (2017), ‘Super-


critical co2 as heat transfer fluid: A review’, Applied thermal engineering
125, 799–810.

Campana, F., Bianchi, M., Branchini, L., De Pascale, A., Peretto, A., Baresi,
M., Fermi, A., Rossetti, N. & Vescovo, R. (2013), ‘Orc waste heat recovery
in european energy intensive industries: Energy and ghg savings’, Energy
Conversion and Management 76, 244–252.

Canale, R. & Chapra, S. (2009), Numerical Methods for Engineers, McGraw-


Hill Science/Engineering/Math.
76 BIBLIOGRAPHY

Carlson, M. D. (2014), Sandia progress on advanced heat exchangers for sco2


brayton cycles., Technical report, Sandia National Lab.(SNL-NM), Albu-
querque, NM (United States).

Cayer, E., Galanis, N. & Nesreddine, H. (2010), ‘Parametric study and op-
timization of a transcritical power cycle using a low temperature source’,
Applied Energy 87(4), 1349–1357.

Cembureau (2019), ‘Key facts & figures’, https://cembureau.eu/about-our-


industry/key-facts-figures/. (Accessed on 07/23/2020).

Conboy, T., Pasch, J. & Fleming, D. (2013), ‘Control of a supercritical co2


recompression brayton cycle demonstration loop’, Journal of engineering
for gas turbines and power 135(11).

Conboy, T., Wright, S., Pasch, J., Fleming, D., Rochau, G. & Fuller, R. (2012),
‘Performance characteristics of an operating supercritical co2 brayton cy-
cle’, Journal of Engineering for Gas Turbines and Power 134(11).

Costiuc, I., Costiuc, L. & Radu, S. (2015), ‘Waste heat recovery using direct
thermodynamic cycle’. [Online; accessed 21. Aug. 2020].

Coutier, J. P. & Farber, E. (1982), ‘Two applications of a numerical approach


of heat transfer process within rock beds’, Solar Energy 29(6), 451–462.

CSI, E. (2017), ‘Development of state of the art techniques in cement manu-


facturing: trying to look ahead’, Eur. Cem. Res. Acad. .

CTCN (2012), ‘Inert anode technology for aluminium smelters’, https://www.


ctc-n.org/technologies/inert-anode-technology-aluminium-smelters. (Ac-
cessed on 07/20/2020).

Cunha, V. M. B., da Silva, M. P., da Costa, W. A., de Oliveira, M. S., Bezerra,


F. W. F., de Melo, A. C., Pinto, R. H. H., Machado, N. T., Araujo, M. E. &
de Carvalho Junior, R. N. (2018), ‘Carbon dioxide use in high-pressure ex-
traction processes’, Carbon Dioxide Chemistry, Capture and Oil Recovery
p. 211.

de Vet, J.-M., Pauer, A., Merkus, E., Baker, P., Gonzalez-Martinez, A. R.,
Kiss-Galfalvi, T., Streicher, G., Rincon-Aznar, A. et al. (2018), ‘Competi-
tiveness of the european cement and lime sectors’, WIFO Studies .
BIBLIOGRAPHY 77

Doğan, A., Bodnarova, B., Hedman, B. A., Avci, F., Feckova, V., Menkova,
V. & Gorbatenko, Y. (2018), Waste heat recovery in turkish cement indus-
try: review of existing installations and assessment of remaining potential,
Technical report, The World Bank.

Dostal, V., Driscoll, M. J. & Hejzlar, P. (2004), A supercritical carbon diox-


ide cycle for next generation nuclear reactors, PhD thesis, Massachusetts
Institute of Technology, Department of Nuclear Engineering . . . .

Dostal, V., Driscoll, M. J., Hejzlar, P. & Todreas, N. E. (2002), A supercritical


co2 gas turbine power cycle for next-generation nuclear reactors, in ‘Inter-
national Conference on Nuclear Engineering’, Vol. 35960, pp. 567–574.

Douvartzides, S. & Karmalis, I. (2016), ‘Working fluid selection for the or-
ganic rankine cycle (ORC) exhaust heat recovery of an internal combustion
engine power plant’, IOP Conference Series: Materials Science and Engi-
neering 161, 012087.

Dyreby, J. J. (2014), Modeling the supercritical carbon dioxide Brayton cycle


with recompression, PhD thesis, The University of Wisconsin-Madison.

ECEEE Discount rates (n.d.), https://www.eceee.org/policy-areas/discount-


rates/. (Accessed on 08/20/2020).

El Samad, T., Amaral Teixeira, J. & Oakey, J. (2020), ‘Investigation of a ra-


dial turbine design for a utility-scale supercritical co2 power cycle’, Applied
Sciences 10(12), 4168.

EU Commission (2015), ‘Eu ets handbook’, Brussels, Belgium .


URL: https:// ec.europa.eu/ clima/ sites/ clima/ files/ docs/ ets_handbook_
en.pdf

EU Commission (2019a), ‘Guidance on determining the allocation at


installation level’, https : / / ec . europa . eu / clima / sites / clima / files / ets /
allowances/docs/p4_gd2_allocation_methodologies_en.pdf. (Accessed on
05/27/2020).

EU Commission (2019b), ‘Waste gases and process emissions sub-installation


for eu ets’, https:// ec.europa.eu/ clima/ sites/ clima/ files/ ets/ allowances/
docs/ p4_ gd8_ waste_ gases_ process_ emissions_ en.pdf. (Accessed on
07/27/2020).
78 BIBLIOGRAPHY

EU Court of Justice (2019), ‘Judgment of the court (fifth chamber) of 20 june


2019 - scheme for greenhouse gas emission allowance trading — natural gas
processing installation’, https://eur-lex.europa.eu/legal-content/EN/TXT/
?uri=CELEX:62017CJ0682. (Accessed on 07/27/2020).

European Union Central bank discount rate - Economy (n.d.), https : / /


www. indexmundi . com / european _ union / central _ bank _ discount _ rate .
html#:~:text=Central%20bank%20discount%20rate%3A%200.25,(31%
20December%202015%20est.)&text=Definition%3A%20This%20entry%
20provides % 20the , meet % 20temporary % 20shortages % 20of % 20funds.
(Accessed on 08/20/2020).

European Union inflation rate | Statista (n.d.), https:// www.statista.com/


statistics / 267908 / inflation - rate - in - eu - and - euro - area/. (Accessed on
08/20/2020).

Eurostat (2020), ‘Eurostat - Data Explorer’, https://appsso.eurostat.ec.europa.


eu/nui/show.do?dataset=nrg_cb_sff%E2%9F%A8=en. [Online; accessed
20. Aug. 2020].

Feher, E. G. (1968), ‘The supercritical thermodynamic power cycle’, Energy


conversion 8(2), 85–90.

Firth, A., Zhang, B. & Yang, A. (2019), ‘Quantification of global waste heat
and its environmental effects’, Applied Energy 235, 1314–1334.

Fleming, D., Holschuh, T., Conboy, T., Rochau, G. & Fuller, R. (2012), Scal-
ing considerations for a multi-megawatt class supercritical co2 brayton cy-
cle and path forward for commercialization, in ‘Turbo Expo: Power for
Land, Sea, and Air’, Vol. 44717, American Society of Mechanical Engi-
neers, pp. 953–960.

Forman, C., Muritala, I. K., Pardemann, R. & Meyer, B. (2016), ‘Estimating


the global waste heat potential’, Renewable and Sustainable Energy Reviews
57, 1568–1579.

Fraunhofer Institute (2009), ‘Methodology for the free allocation of emis-


sion allowances in the eu for the aluminium industry’, https://ec.europa.eu/
clima/sites/clima/files/ets/allowances/docs/bm_study-aluminium_en.pdf.
(Accessed on 07/21/2020).

Freeman, J., Hellgardt, K. & Markides, C. N. (2015), ‘An assessment of solar-


powered organic rankine cycle systems for combined heating and power in
BIBLIOGRAPHY 79

uk domestic applications’, Applied Energy 138, 605 – 620.


URL: https:// doi.org/ 10.1016/ j.apenergy.2014.10.035

Ge, Y., Li, L., Luo, X. & Tassou, S. (2018), ‘Performance evaluation of a low-
grade power generation system with co2 transcritical power cycles’, Applied
Energy 227, 220–230.

Gerdes, K., Summers, W. M. & Wimer, J. (2011), Quality guidelines for en-
ergy system studies: Cost estimation methodology for netl assessments of
power plant performance, Technical report, National Energy Technology
Laboratory (NETL), Pittsburgh, PA, Morgantown, WV . . . .

Gokhshtein, D., KOZOREZ, A. & DEKHTYAREV, V. (1971), ‘Future de-


signs of thermal power-stations operating on carbon-dioxide’, Thermal En-
gineering 18(4), 54–+.

Gokhshtein, D. & Verkhivker, G. (1969), ‘Use of carbon dioxide as a heat car-


rier and working substance in atomic power stations’, Soviet Atomic Energy
26(4), 430–432.

Gómez-Aláez, S. L., Brizzi, V., Alfani, D., Silva, P., Giostri, A. & Astolfi,
M. (2017), ‘Off-design study of a waste heat recovery orc module in gas
pipelines recompression station’, Energy Procedia 129, 567–574.

Haenel, R., Stegena, L. & Rybach, L. (2012), Handbook of terrestrial heat-


flow density determination: with guidelines and recommendations of the
International Heat Flow Commission, Vol. 4, Springer Science & Business
Media.

Hammond, G. & Norman, J. (2014), ‘Heat recovery opportunities in uk indus-


try’, Applied Energy 116, 387–397.

Holman, J. (2018), Heat Transfer (Int’l Ed), McGraw-Hill Education / Asia.

IEA (2019), ‘Global CO2 emissions in 2019 – Analysis - IEA’, https://www.


iea.org/articles/global-co2-emissions-in-2019. [Online; accessed 20. Aug.
2020].

Industrial electricity prices across Europe 2018 | Statista (n.d.), https://www.


statista.com/statistics/267068/industrial-electricity-prices-in-europe/#:~:
text=In%202018%2C%20industrial%20end%2Dusers, hours%20paid%
208.7%20euro%20cents. (Accessed on 07/09/2020).
80 BIBLIOGRAPHY

Industrial production statistics - Statistics Explained (n.d.), https : / / ec .


europa.eu/eurostat/statistics-explained/index.php/Industrial_production_
statistics#Industrial_production_by_country. (Accessed on 07/17/2020).

Johnson, G. A., McDowell, M. W., O’Connor, G. M., Sonwane, C. G. & Sub-


baraman, G. (2013), ‘Supercritical CO2 Cycle Development at Pratt and
Whitney Rocketdyne’, American Society of Mechanical Engineers Digital
Collection pp. 1015–1024.

Johnson, I., Choate, W. T. & Davidson, A. (2008), Waste heat recovery. tech-
nology and opportunities in us industry, Technical report, BCS, Inc., Laurel,
MD (United States).

Jouhara, H., Khordehgah, N., Almahmoud, S., Delpech, B., Chauhan, A. &
Tassou, S. A. (2018), ‘Waste heat recovery technologies and applications’,
Thermal Science and Engineering Progress 6, 268–289.

Karaboga, D. & Basturk, B. (2007), Artificial bee colony (abc) optimiza-


tion algorithm for solving constrained optimization problems, in P. Melin,
O. Castillo, L. T. Aguilar, J. Kacprzyk & W. Pedrycz, eds, ‘Foundations
of Fuzzy Logic and Soft Computing’, Springer Berlin Heidelberg, Berlin,
Heidelberg, pp. 789–798.

Kenton, W. (2019), ‘Discounted Payback Period Definition’, Investopedia .


URL: https : / / www. investopedia . com / terms / d / discounted - payback -
period.asp

Kenton, W. (2020), ‘Net Present Value (NPV)’, Investopedia .


URL: https:// www.investopedia.com/ terms/ n/ npv.asp

Kizilkan, O. (2020), ‘Performance assessment of steam rankine cycle and sco2


brayton cycle for waste heat recovery in a cement plant: A comparative
study for supercritical fluids’, International Journal of Energy Research .

KLM (2011), ‘Process design of air cooled heat exchangers (air cool-
ers) (project standards and specifications)’, https:// www.klmtechgroup.
com/PDF/ess/PROJECT_STANDARDS_AND_SPECIFICATIONS_air_
cooled_exchangers_Rev01.pdf. (Accessed on 07/15/2020).

Koelet, P. C. & Gray, T. B. (2017), Industrial refrigeration: principles, design


and applications, Macmillan International Higher Education.
BIBLIOGRAPHY 81

Kosmadakis, G. (2019), ‘Estimating the potential of industrial (high-


temperature) heat pumps for exploiting waste heat in eu industries’, Applied
Thermal Engineering 156, 287–298.

Le Moullec, Y. (2013), ‘Conceptual study of a high efficiency coal-fired power


plant with co2 capture using a supercritical co2 brayton cycle’, Energy
49, 32–46.

Lead (n.d.), ‘Sources of co2’, CARBON DIOXIDE CAPTURE AND STORAGE


p. 75.

Li, L., Ge, Y., Luo, X. & Tassou, S. (2018), ‘Experimental analysis and com-
parison between co2 transcritical power cycles and r245fa organic rankine
cycles for low-grade heat power generations’, Applied Thermal Engineering
136, 708–717.

Li, X., Shu, G., Tian, H., Shi, L., Huang, G., Chen, T. & Liu, P. (2017), ‘Pre-
liminary tests on dynamic characteristics of a co2 transcritical power cycle
using an expansion valve in engine waste heat recovery’, Energy 140, 696–
707.

Linhares, C., Niceas, F., Bacelar, R., Campelo, H., Silva, M. A., Araujo, J. &
Reis, R. (2016), Energy efficiency improvement in anode baking furnaces,
in ‘Light Metals 2013’, Springer, pp. 1151–1154.

Liu, Y., Wang, Y. & Huang, D. (2019), ‘Supercritical co2 brayton cycle: A
state-of-the-art review’, Energy 189, 115900.

Lu, H., Price, L. & Zhang, Q. (2016), ‘Capturing the invisible resource: Anal-
ysis of waste heat potential in chinese industry’, Applied energy 161, 497–
511.

Ma, X., Cao, Q. & Cui, Z. (2019), ‘Optimization design of the grate cooler
based on the power flow method and genetic algorithms’, Journal of Ther-
mal Science pp. 1–10.

Manente, G. & Lazzaretto, A. (2014), ‘Innovative biomass to power conver-


sion systems based on cascaded supercritical co2 brayton cycles’, Biomass
and Bioenergy 69, 155–168.

Marti, J., Geissbühler, L., Becattini, V., Haselbacher, A. & Steinfeld, A.


(2018), ‘Constrained multi-objective optimization of thermocline packed-
bed thermal-energy storage’, Applied Energy 216, 694–708.
82 BIBLIOGRAPHY

Martínez, P. E. (1992), Termodinámica básica y aplicada, Escuela Universi-


taria Politécnica de Albacete, Servicio de Publicaciones.

McBrien, M., Serrenho, A. C. & Allwood, J. M. (2016), ‘Potential for en-


ergy savings by heat recovery in an integrated steel supply chain’, Applied
Thermal Engineering 103, 592–606.

McKenna, R. C. & Norman, J. B. (2010), ‘Spatial modelling of industrial heat


loads and recovery potentials in the uk’, Energy Policy 38(10), 5878–5891.

McTigue, J. D. P., Farres-Antunez, P., Ellingwood, K., Neises, T. W. & White,


A. (2019), Pumped thermal electricity storage with supercritical co2 cy-
cles and solar heat input, Technical report, National Renewable Energy
Lab.(NREL), Golden, CO (United States).

Mercangöz, M., Hemrle, J., Kaufmann, L., Z’Graggen, A. & Ohler, C.


(2012), ‘Electrothermal energy storage with transcritical co2 cycles’, En-
ergy 45(1), 407–415.

Milewski, J. & Krasucki, J. (2018), ‘Comparison of orc and kalina cycles for
waste heat recovery in the steel industry’, Journal of power technologies
97(4), 302–307.

Mlcak, H. A. (n.d.), Pe, an introduction to the kalina cycle, pwr-vol. 30, in


‘Proceedings of the International Joint Power Generation Conference Edi-
tors: L Kielasa, and GE Weed Book’, number H01077-1996.

Moisseytsev, A., Sienicki, J. J. & Lv, Q. (n.d.), ‘Dry air cooler modeling for
supercritical carbon dioxide brayton cycle analysis’.

Morandin, M., Mercangöz, M., Hemrle, J., Maréchal, F. & Favrat, D. (2013),
‘Thermoeconomic design optimization of a thermo-electric energy storage
system based on transcritical co2 cycles’, Energy 58, 571–587.

Moya, J., Pardo, N. & Mercier, A. (2010), ‘Energy efficiency and co2 emis-
sions: prospective scenarios for the cement industry’, JRC Scientific and
Technical Report No. EUR 24592.

Musgrove, G., Pittaway, C., Vollnogle, E. & Chordia, L. (2014), Tutorial: heat
exchangers for supercritical co2 power cycle applications, in ‘ASME Turbo
Expo’, pp. 24–7.
BIBLIOGRAPHY 83

Naik-Dhungel, N. (2012), ‘Waste heat to power systems’, https://www.epa.


gov/ sites/ production/ files/ 2015- 07/ documents/ waste_ heat_ to_ power_
systems.pdf. [Online; accessed 18. Aug. 2020].

Ndlovu, S., Simate, G. S. & Matinde, E. (2017), Waste production and utiliza-
tion in the metal extraction industry, CRC Press.

Nellis, G. & Klein, S. (2012), Heat Transfer, Cambridge University Press.

Nemati, A., Nami, H., Ranjbar, F. & Yari, M. (2017), ‘A comparative thermo-
dynamic analysis of orc and kalina cycles for waste heat recovery: A case
study for cgam cogeneration system’, Case Studies in Thermal Engineering
9, 1 – 13.
URL: https:// doi.org/ 10.1016/ j.csite.2016.11.003

Nowicki, C. & Gosselin, L. (2012), ‘An overview of opportunities for waste


heat recovery and thermal integration in the primary aluminum industry’,
Jom 64(8), 990–996.

Pan, L., Li, B., Wei, X. & Li, T. (2016), ‘Experimental investigation on the
co2 transcritical power cycle’, Energy 95, 247–254.

Papapetrou, M., Kosmadakis, G., Cipollina, A., La Commare, U. & Micale,


G. (2018), ‘Industrial waste heat: Estimation of the technically available
resource in the eu per industrial sector, temperature level and country’, Ap-
plied Thermal Engineering 138, 207–216.

Paspaliaris, I., Panias, D., Amanatidis, A., Mordini, J., Werner, D., Panou, G.
& Ballas, D. (1999), ‘Precipitation and calcination of monohydrate alumina
from the bayer process liquors’, Eurothen 99, 532–547.

Patel, V. K., Savsani, V. J. & Tawhid, M. A. (2019), Thermal System Opti-


mization - A Population-Based Metaheuristic Approach | Vivek K. Patel |
Springer, Springer International Publishing.

Pecnik, R. & Colonna, P. (2011), Accurate cfd analysis of a radial compressor


operating with supercritical co2, in ‘Supercritical CO2 Power Cycle Sym-
posium, Boulder, Colorado, USA’.

Pecnik, R., Rinaldi, E. & Colonna, P. (2012), ‘Computational fluid dynamics


of a radial compressor operating with supercritical co2’, Journal of Engi-
neering for Gas Turbines and Power 134(12).
84 BIBLIOGRAPHY

Persichilli, M., Held, T., Hostler, S., Zdankiewicz, E. & Klapp, D. (2011),
‘Transforming waste heat to power through development of a co2-based-
power cycle’, Electric Power Expo 10.

Petr, V. & Kolovratnik, M. (1997), ‘A study on application of a closed cy-


cle co2 gas turbine in power engineering’, Department of Fluid Dynamics
and Power Engineering, Division of Power Engineering, Czech Technical
University in Prague, Prague, Czech Republic, Report No. Z-523/97 .

Petr, V., Kolovratnik, M. & Hanzal, V. (1999), ‘On the use of co2 gas turbine in
power engineering’, Department of Fluid Dynamics and Power Engineer-
ing, Division of Power Engineering, Czech Technical University, Prague,
Czech Republic, Report No. Z-530/99 .

Pfost, H. & Seitz, K. (1971), Characteristics of an installation employing a co


sub2 gas turbine cycle running at over-critical pressure., Technical report,
Firma Kraftwerk Union, Erlangen, Ger.

Qin, S. & Chang, S. (2017), ‘Modeling, thermodynamic and techno-economic


analysis of coke production process with waste heat recovery’, Energy
141, 435–450.

Qingcai, Z., Jingli, Z., Xingli, M. & Lihai, W. (2012), Energy saving technolo-
gies for anode manufacturing, in ‘Light Metals 2012’, Springer, pp. 1201–
1204.

Quoilin, S., Broek, M. V. D., Declaye, S., Dewallef, P. & Lemort, V. (2013),
‘Techno-economic survey of organic rankine cycle (orc) systems’, Renew-
able and Sustainable Energy Reviews 22, 168 – 186.
URL: https:// doi.org/ 10.1016/ j.rser.2013.01.028

Rentizelas, A., Karellas, S., Kakaras, E. & Tatsiopoulos, I. (2009), ‘Compar-


ative techno-economic analysis of orc and gasification for bioenergy appli-
cations’, Energy Conversion and Management 50, 674–681.

Rinaldi, E., Pecnik, R. & Colonna, P. (2013), Steady state cfd investigation
of a radial compressor operating with supercritical co2, in ‘Turbo Expo:
Power for Land, Sea, and Air’, Vol. 55294, American Society of Mechanical
Engineers, p. V008T34A008.

Rinaldi, E., Pecnik, R. & Colonna, P. (2014), Numerical computation of


the performance map of a supercritical co2 radial compressor by means
BIBLIOGRAPHY 85

of three-dimensional cfd simulations, in ‘Turbo Expo: Power for Land,


Sea, and Air’, Vol. 45660, American Society of Mechanical Engineers,
p. V03BT36A017.

Rinaldi, E., Pecnik, R. & Colonna, P. (2015), ‘Computational fluid dynamic


simulation of a supercritical co2 compressor performance map’, Journal of
Engineering for Gas Turbines and Power 137(7).

Rogdakis, E. & Lolos, P. (2015), ‘Kalina cycles for power generation’, Hand-
book of Clean Energy Systems pp. 1–25.

ROSSIN, D. (2013), ‘Modellizzazione di un impianto di trigenerazione appli-


cato alla centrale di ricompressione gas di poggio renatico (fe)’.

Saadaoui, H. (2020), ‘Industrial waste heat and waste heat recovery systems’,
https://haemers-technologies.com/industrial-waste-heat-recovery-systems.
[Online; accessed 20. Aug. 2020].

Saleh, B., Koglbauer, G., Wendland, M. & Fischer, J. (2007), ‘Working fluids
for low-temperature organic rankine cycles’, Energy 32(7), 1210 – 1221.
URL: https:// doi.org/ 10.1016/ j.energy.2006.07.001

Santini, L., Accornero, C. & Cioncolini, A. (2016), ‘On the adoption of carbon
dioxide thermodynamic cycles for nuclear power conversion: A case study
applied to mochovce 3 nuclear power plant’, Applied Energy 181, 446–463.

Satyendra (2015), ‘Coke oven gas, its characteristics and safety requirements –
ispatguru’, https://www.ispatguru.com/coke-oven-gas-its-characteristics-
and-safety-requirements/. (Accessed on 07/23/2020).

Scarsella, A. (n.d.), ‘Generation 5 calciner’, https : / / www. outotec . com /


products - and - services / newsletters / aluminum - insights / generation - 5 -
calciner/. (Accessed on 07/21/2020).

Schmidt, D. (1970), ‘Co2 gas turbine power plant’. US Patent 3,503,208.

Schumann, T. E. (1929), ‘Heat transfer: a liquid flowing through a porous


prism’, Journal of the Franklin Institute 208(3), 405–416.

sCO2-flex (2019), https://www.sco2-flex.eu/. (Accessed on 08/1/2020).

Seo, H., Cha, J. E., Kim, J., Sah, I. & Kim, Y.-W. (2020), ‘Design and perfor-
mance analysis of a supercritical carbon dioxide heat exchanger’, Applied
Sciences 10(13), 4545.
86 BIBLIOGRAPHY

Shi, L., Shu, G., Tian, H., Huang, G., Chang, L., Chen, T. & Li, X. (2017),
‘Ideal point design and operation of co2-based transcritical rankine cycle
(ctrc) system based on high utilization of engine’s waste heats’, Energies
10(11), 1692.

Shi, L., Shu, G., Tian, H., Huang, G., Chen, T., Li, X. & Li, D. (2017), ‘Ex-
perimental comparison between four co2-based transcritical rankine cycle
(ctrc) systems for engine waste heat recovery’, Energy Conversion and Man-
agement 150, 159–171.

Shiferaw, D. (2017), ‘Heat exchangers in sco2 power conversion cycles’, http://


sco2symposium.com/papers2016/Tutorials/HeatExchanger.pdf. (Accessed
on 08/20/2020).

Singh, S. & Dasgupta, M. (2017), ‘Co2 heat pump for waste heat recovery
and utilization in dairy industry with ammonia based refrigeration’, Inter-
national Journal of Refrigeration 78, 108–120.

Song, J., Loo, P., Teo, J. & Markides, C. N. (2020), ‘Thermo-economic opti-
mization of organic rankine cycle (orc) systems for geothermal power gen-
eration: A comparative study of system configurations’, Frontiers in Energy
Research 8, 6.
URL: https:// www.frontiersin.org/ article/ 10.3389/ fenrg.2020.00006

Sørhuus, A. & Wedde, G. (2016), Pot gas heat recovery and emission control,
in ‘Essential Readings in Light Metals’, Springer, pp. 987–992.

Span, R. & Wagner, W. (1996), ‘A new equation of state for carbon dioxide
covering the fluid region from the triple-point temperature to 1100 k at pres-
sures up to 800 mpa’, Journal of Physical and Chemical Reference Data
25(6), 1509–1596.
URL: https:// doi.org/ 10.1063/ 1.555991

Spelling, J. (2013), Hybrid solar gas-turbine power plants: a thermoeconomic


analysis, PhD thesis, KTH Royal Institute of Technology.

Standard, A. & Edition, F. (2002), ‘Air-cooled heat exchangers for general


refinery service’.

Strub, R. & Frieder, A. (1970), Paper 5: High pressure indirect co2 closed-
cycle gas turbines, in ‘Nuclear gas turbines’, Thomas Telford Publishing,
pp. 51–61.
BIBLIOGRAPHY 87

Swedish Energy Agency (2019), ‘Emissions trading in the eu’, http : / /


www.energimyndigheten.se/en/sustainability/eu-ets---implementation-in-
sweden/about-emissions-trading/emissions-trading-in-the-eu/. (Accessed
on 07/27/2020).

SWEPCO Fact Sheet (n.d.), https:// www.swepco.com/ global/ utilities/ lib/


docs/info/facts/factsheets/SWEPCOFactSheetJuly2017.pdf. (Accessed on
02/15/2020).

Tajik, A. R., Shamim, T., Al-Rub, R. K. A. & Zaidani, M. (2017), Performance


analysis of a horizontal anode baking furnace for aluminum production, in
‘ICTEA: International Conference on Thermal Engineering’, Vol. 2017.

The International Aluminium Institute (n.d.), ‘Mining and refining – energy ef-
ficiency’, http://bauxite.world-aluminium.org/refining/energy-efficiency/.
(Accessed on 07/21/2020).

Tiskatine, (2017), ‘Identification of suitable storage materials for solar thermal


power plant using selection methodology’, Applied Thermal Engineering
117, 591–608.

Tiskatinee, . R., Oaddi, R., El Cadi, R. A., Bazgaou, A., Bouirden, L.,
Aharoune, A. & Ihlal, A. (2017), ‘Suitability and characteristics of rocks
for sensible heat storage in csp plants’, Solar Energy Materials and Solar
Cells 169, 245–257.

Tran, K., Harp, G., Sigurbjörnsson, O., Bergins, C. & Buddenberg, T. (2016),
‘Carbon recycling for converting coke oven gas to methanol for the reduction
of carbon dioxide at steel mills’, MefCO2: Auderghem, Belgium .

Trevisan, S., Guédez, R. & Laumert, B. (2019), Supercritical co2 brayton


power cycle for csp with packed bed tes integration and cost benchmark
evaluation, in ‘ASME Power Conference’, Vol. 59100, American Society of
Mechanical Engineers, p. V001T06A010.

Vesely, L., Dostal, V. & Hajek, P. (2014), Design of experimental loop


with supercritical carbon dioxide, in ‘International Conference on Nuclear
Engineering’, Vol. 45936, American Society of Mechanical Engineers,
p. V003T05A023.

Vinnemeier, P., Wirsum, M., Malpiece, D. & Bove, R. (2016), ‘Integration of


heat pumps into thermal plants for creation of large-scale electricity storage
capacities’, Applied Energy 184, 506–522.
88 BIBLIOGRAPHY

Wang, J., Huang, Y., Zang, J. & Liu, G. (2014), Research activities on su-
percritical carbon dioxide power conversion technology in china, in ‘Turbo
Expo: Power for Land, Sea, and Air’, Vol. 45660, American Society of Me-
chanical Engineers, p. V03BT36A009.

Wang, X. & Dai, Y. (2016), ‘Exergoeconomic analysis of utilizing the tran-


scritical co2 cycle and the orc for a recompression supercritical co2 cycle
waste heat recovery: A comparative study’, Applied Energy 170, 193–207.

Watzel, G. (1971), ‘Is the co 2 gas turbine cycle an economical proposition for
fast sodium cooled breeder reactors’, Brennstoff-Waerme-Kraft 23(9), 395–
400.

Weiland, N. (2019), Techno-economics and cost modeling for state-of-the-art


sco2 components, Technical report, NETL.

Weiland, N., Lance, B. & Pidaparti, S. (2019), sco2 power cycle component
cost correlations from doe data spanning multiple scales and applications
(final), Technical report, NETL.

Weiland, N. & Shelton, W. (2016), ‘Systems analyses of direct power ex-


traction (dpe) and advanced ultra-supercritical (ausc) power plants’, https:
// netl.doe.gov/ sites/ default/ files/ event- proceedings/ 2016/ crosscutting-
ree/Systems-Analysis_AUSC-and-DPE-Assessment.pdf. (Accessed on
07/15/2020).

Weiland, N., Shelton, W., Shultz, T., White, C. W. & Gray, D. (2017), Per-
formance and cost assessment of a coal gasification power plant integrated
with a direct-fired sco2 brayton cycle, Technical report, National Energy
Technology Laboratory (NETL), Pittsburgh, PA, Morgantown, WV . . . .

White, C., Gray, D., Plunkett, J., Shelton, W., Weiland, N. & Shultz, T. (2017),
‘Techno-economic evaluation of utility-scale power plants based on the in-
direct sco2 brayton cycle’. [Online; accessed 20. Aug. 2020].

White, C. W., Shelton, W., Weiland, N. & Shultz, T. R. (2018), sco2 cycle as an
efficiency improvement opportunity for air-fired coal combustion, in ‘The
6th International Supercritical CO2 Power Cycles Symposium’, pp. 27–29.

World Aluminium (2020), ‘Primary aluminium production’, http:// www.


world-aluminium.org/statistics/. (Accessed on 07/20/2020).
BIBLIOGRAPHY 89

Wortler, M., Schuler, F., Voigt, N., Schmidt, T., Dahlmann, P., Lüngen, H.
& Ghenda, J. (2013), ‘Steel’s contribution to a low carbon europe 2050’,
Technical and economic analysis of the sectors CO2 abatement potential,
Boston .

Wright, S. A., Davidson, C. S. & Scammell, W. O. (2016), Thermo-economic


analysis of four sco2 waste heat recovery power systems, in ‘Fifth Interna-
tional SCO2 Symposium, San Antonio, TX, Mar’, pp. 28–31.

Yasuyoshi, K., Takeshi, N. & Yoshio, Y. (2001), A carbon dioxide partial con-
densation direct cycle for advanced gas cooled fast and thermal reactors,
in ‘Proceedings of Global 2001 international conference on:“back-end of
the fuel cycle: from research to solutions”, INIS-FR-1118, Paris, France’,
pp. 9–13.

Yingying, L., Laishi, L., Xinqin, L. & Ruisheng, B. (2015), Recovering waste-
heat and water from alumina calciner gas, in ‘Light Metals 2015’, Springer,
pp. 69–71.

Yu, M., Gudjonsdottir, M. S., Valdimarsson, P. & Saevarsdottir, G. (2018),


Waste heat recovery from aluminum production, in ‘TMS Annual Meeting
& Exhibition’, Springer, pp. 165–178.

Zhang, H., Zhao, H., Deng, Q. & Feng, Z. (2015), Aerothermodynamic de-
sign and numerical investigation of supercritical carbon dioxide turbine, in
‘Turbo Expo: Power for Land, Sea, and Air’, Vol. 56802, American Society
of Mechanical Engineers, p. V009T36A007.

Zhang, S. (2019), ‘Pilot study on ascension-pipe heat exchanger used for waste
heat recovery of coke oven gas’, Energy Procedia 158, 26–31.

Zhang, X.-R. & Yamaguchi, H. (2011), ‘An experimental investigation on


characteristics of supercritical co2-based solar rankine system’, Interna-
tional journal of energy research 35(13), 1168–1178.

Zhang, X.-R., Yamaguchi, H., Fujima, K., Enomoto, M. & Sawada, N. (2005),
‘A feasibility study of co2-based rankine cycle powered by solar energy’,
JSME International Journal Series B Fluids and Thermal Engineering
48(3), 540–547.

Zhang, X.-R., Yamaguchi, H. & Uneno, D. (2007), ‘Experimental study on the


performance of solar rankine system using supercritical co2’, Renewable
Energy 32(15), 2617–2628.
90 BIBLIOGRAPHY

Zhang, X.-R., Yamaguchi, H., Uneno, D., Fujima, K., Enomoto, M. & Sawada,
N. (2006), ‘Analysis of a novel solar energy-powered rankine cycle for com-
bined power and heat generation using supercritical carbon dioxide’, Renew-
able Energy 31(12), 1839–1854.

Zhang, X. & Yamaguchi, H. (2008), ‘An experimental study on evacuated


tube solar collector using supercritical co2’, Applied Thermal Engineering
28(10), 1225–1233.

Zhao, R. (2015), ‘Analysis, simulation and optimization of ventilation of alu-


minum smelting cells and potrooms for waste heat recovery’.
BIBLIOGRAPHY 91

Appendix
Appendix A. Cost Functions
Turbomachinery Costs equations
CC =CRef . W exp
CT =fT · CRef ·W exp
(
1 if T IT < 550 [◦ C]
fT =
1 + d · (T IT [◦ C] − 550)2 if T IT ≥ 550 [◦ C]

d is 1.11.10−4 for axial turbines and 1.137.10−5 for radial turbines


CGen =CRef . W exp
CGB =CRef . W exp
CM ot =CRef . W exp
Where: W is the power capacity of the component in MW.
TIT is the Turbine Inlet Temperature.
CT is the cheapest of the two turbine types
CM ot is the cheapest of the valid Motor types
Explosion Proof Motor model limited between 0.00075-2.8MW
Synchronous Motor model limited between 0.15 to 15 MW
Open-Drip Proof Motor model limited between 0.00075-37 MW

Heat Exchangers
CHEX =CRef . (U A)exp
Where: U [W/m2 K] is the overall heat transfer coefficient, multiplied by the
heat transfer area, A of the heat exchanger.
CRecuperator =fT · CRef · (U A)exp
(
1 if Tmax < 550 [◦ C]
fT =
1 + c · (Tmax [◦ C] − 550) if Tmax ≥ 550 [◦ C]

Cost of Piping:

Cpip =CT . 0.10 Depends on the temperatures in the system.


92 BIBLIOGRAPHY

Table 6.1: Components’ Reference Prices and Exponents


Component Cref ($) Exponent Reference
Axial Turbine 182,600 0.5561 (Weiland et al. 2019)
Radial Turbine 406,200 0.8 (Weiland et al. 2019)
Compressor 1,230,000 0.3992 (Weiland et al. 2019)
Generator 108,900 0.5463 (Weiland et al. 2019)
Gearbox 177,200 0.2434 (Weiland et al. 2019)
Explosion Proof Motor 131,40 0.5611 (Weiland et al. 2019)
Synchronous Motor 211,400 0.6227 (Weiland et al. 2019)
Drip Proof Motor 399,400 0.6062 (Weiland et al. 2019)
Waste Heat Exchanger 5 1 (Wright et al. 2016)
Recuperator 49.45 0.7544 (Weiland et al. 2019)
Air Cooler 32.88 0.75 (Weiland et al. 2019)

Table 6.2: Storage Component Costs


Specific Cost, Ci Value Unit Reference
High Temperature Insulation 4,269 [$/m3] (Marti et al. 2018)
Low Temperature Insulation 616 [$/m3] (Marti et al. 2018)
253 MA Stainless Steel 42,354 [$/m3] (Trevisan et al. 2019)
Natural Rocks 66 [$/m3] (Marti et al. 2018)
Foundations 1210 [$/m2] (Marti et al. 2018)

Cost of Storage:
X πD2
CT ES = Ci .Vi + Cf ound .NT ES ( )
4
Where:
CT ES is the cost of the storage
Ci is the cost of each component
Vi is the volume of each component
NT ES is the number of storage systems
D is the diameter of the storage system

Balance of Plant: This includes miscellaneous balance of plant (38 $/kWe ),


Instrumentation and Control (43 $/kWe ) along with Site improvements (30
$/kWe ) (White et al. 2018).

CBOP =(38+43+30) . (Pe )


BIBLIOGRAPHY 93

Table 6.3: Labor and Material Costs’ share of Component Costs (Weiland
2019)
Component Labor Extra Cost Materials Extra Cost
Turbine 12% 8%
Compressor 12% 8%
Generator 12% 8%
Gearbox 12% 8%
Motor 12% 8%
Waste Heat Exchanger N/A N/A
Recuperator 3% 2%
Air Cooler 12% 8%

Labor & Material: Labor and Material costs vary from location to another.
Therefore, they were not included in some of the above mentioned prices. It
might appear that material costs are included in equipment cost. However,
sCO2 component costs might be more expensive due to upgraded materials
needed for the higher operating temperatures. Their estimates are given in Ta-
ble 6.3 (Weiland 2019).

Direct Capital Cost:

CAP EXDirect =CsCO2 + CT ES + CBOP + CLab + CM at


Indirect Capital Cost:

CAP EXIndirect = CEP C + CCont

Engineering, Procurement and Construction:


CEP C =CAP EXDirect . 0.09
(Gerdes et al. 2011)

Contingencies:

CCont =CAP EXDirect . 0.3


(Gerdes et al. 2011)

CAPEX=CAP EXDirect +CAP EXIndirect


Operation & Maintenance:

OPEX= CT ax + CLabor + CV ariableOM


94 BIBLIOGRAPHY

CT ax =0.02 . CAPEX (Weiland et al. 2017)

CLabor = 63 . (W/1000) (Weiland et al. 2017) CV ariableOM = 9.70 . (W/1000)


(Weiland et al. 2017)
Decommissioning:

CDecom = 0.05 . (CAP EXDirect +CostCong ) (Spelling 2013)

Appendix B. Sensitivity Analysis

Figure 6.1: Sensitivity Analysis of Change in Compressor Cost


BIBLIOGRAPHY 95

Figure 6.2: Sensitivity Analysis of Change in Turbine Cost

Figure 6.3: Sensitivity Analysis of Change in Recuperator Cost

Figure 6.4: Sensitivity Analysis of Change in Condenser Cost


96 BIBLIOGRAPHY

Figure 6.5: Sensitivity Analysis of Change in Discount Rate

Figure 6.6: Sensitivity Analysis of Change in DownTime


BIBLIOGRAPHY 97

Figure 6.7: Sensitivity Analysis of Change in Waste Heat Exchanger Price

Appendix C. Cement Survey Answers


98 BIBLIOGRAPHY

Figure 6.8: Some Survey Answers as Received from the Cement Industry

You might also like