Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Techno Economic Study of The Calcium Looping Process For CO2 Capture From Cement and Biomass Power Plants

Download as pdf or txt
Download as pdf or txt
You are on page 1of 240

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/284033028

Techno Economic Study of the Calcium Looping Process for CO2 Capture from
Cement and Biomass Power Plants

Thesis · November 2014

CITATIONS READS

8 805

1 author:

Dursun Can Ozcan


The Scientific & Technological Research Council of Türkiye
15 PUBLICATIONS   291 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Dursun Can Ozcan on 18 November 2015.

The user has requested enhancement of the downloaded file.


This thesis has been submitted in fulfilment of the requirements for a postgraduate degree
(e.g. PhD, MPhil, DClinPsychol) at the University of Edinburgh. Please note the following
terms and conditions of use:

• This work is protected by copyright and other intellectual property rights, which are
retained by the thesis author, unless otherwise stated.
• A copy can be downloaded for personal non-commercial research or study, without
prior permission or charge.
• This thesis cannot be reproduced or quoted extensively from without first obtaining
permission in writing from the author.
• The content must not be changed in any way or sold commercially in any format or
medium without the formal permission of the author.
• When referring to this work, full bibliographic details including the author, title,
awarding institution and date of the thesis must be given.
Techno-Economic Study of the Calcium
Looping Process for CO2 Capture from
Cement and Biomass Power Plants

Dursun Can Ozcan

Thesis submitted for the degree of Doctor of Philosophy to


The University of Edinburgh

The School of Engineering, The University of Edinburgh

2014
Declaration

I certify that the work presented in this dissertation was composed by me and that the
work contained herein is my own except where explicitly stated otherwise in the text.
Furthermore, I declare that this work has not been submitted for any other degree or
professional qualification except as specified.

The work presented in this dissertation has contributed to 5 scientific papers,


including the journals of Greenhouse Gas Control and Industrial & Engineering
Chemistry Research, has been presented in 8 international conferences, including the
AIChE Annual Meeting and the IEAGHG HTSLCN Meeting, and has received 2
awards from Honeywell and Association of British and Turkish Academics. Further
details are available in Appendix E.

Dursun Can Ozcan


Lay Summary

The greenhouse effect is a natural process where greenhouse gases trap heat from the
Sun in the atmosphere, warming the surface of the Earth. The greenhouse gases
include water vapour, carbon dioxide (CO2), methane and nitrous oxide and naturally
exist in the atmosphere. This process maintains the Earth’s temperature at a safe
level to support life. Nevertheless, some human activities particularly burning fossil
fuels and clearing of forests intensify the greenhouse effect because more greenhouse
gases are released into the atmosphere. For this reason, the amount of heat absorbed
by the greenhouse gases increases, which results in a rise in global surface
temperatures and sea levels due to the melting of land ice. CO2 is the most important
greenhouse gas acting in climate change. It is expected that the anthropogenic CO2
emissions will continue to increase if fossil fuels remain as the major source of
energy production.

Carbon capture and storage (CCS) is a promising and an emerging way of reducing
CO2 emissions from large point sources, especially fossil-fuelled power stations and
industrial processes such as cement plants. In addition, the use of sustainably-grown
biomass as a renewable energy source in combustion systems is linked to zero net
CO2 emissions. We have developed comprehensive process flowsheets where fully
integrated CCS systems are fitted to cement and biomass-fired power plants.
Calcium looping (Ca-looping) process where CO2 reacts with calcium oxide (CaO)
and is thereby separated from flue gas has been selected as the main option for this
purpose. The Ca-looping process uses relatively cheap and abundant CaO-based
sorbents and is currently being piloted at the ~2 MWth scale. We have evaluated the
techno-economic performance of the Ca-looping process as well as several
alternative CCS technologies. The findings of this dissertation contribute to the
development of efficient and cost effective CCS systems applicable to cement and
biomass-fired power plants.
Abstract

The first detailed systematic investigation of a cement plant with various carbon
capture technologies has been performed. The calcium looping (Ca-looping) process
has emerged as a leading option for this purpose, since this process applied to a
cement plant provides an opportunity to use the CaO purge for clinker production.
The Ca-looping process is comprised of two interconnected reactors where the
carbonator captures CO2 from flue gases and the calciner regenerates the CaCO3 into
CaO by oxy-combustion. Fully integrated process flowsheets have been developed
and simulated in UniSim Design Suite from Honeywell. The detailed carbonator
model has been implemented using Matlab and incorporated into UniSim to provide
a full flowsheet simulation for an exemplary dry-feed cement plant as a user-defined
operation. The base cement plant simulation was also modified to integrate three
different carbon capture processes: membrane; indirect calcination; and amine-
scrubbing. Furthermore, an advanced configuration of Ca-looping process has been
investigated where the energy intensive air separation unit was replaced with a
chemical looping combustion (CLC) cycle. Each case has been optimised to
minimise its energy consumption and compared in terms of levelised cost of cement
and its resulting cost of CO2 avoided at the same CO2 avoidance rate.

The proposed integration of the Ca-looping process is capable of achieving over 90%
CO2 avoidance with additional fuel consumption of 2.5 to 3.0 GJth/ton CO2 avoided.
By using an advanced configuration of the Ca-looping process with a CLC cycle, the
additional fuel consumption can be reduced to 1.7 GJth/ton CO2 avoided, but the cost
of the oxygen carrier is the major concern for this system. Among the other CO2
capture options, the membrane process is a promising alternative for the Ca-looping
process since it has a potential of achieving the target CO2 avoidance rate and purity
requiring lower energy consumption. The indirect calcination process provides
moderate levels of CO2 avoidance (up to 56%) without a need of an external capture
process whereas the integration of the amine process in a cement plant is challenging
as a result of the requirement of steam for solvent regeneration.
Furthermore, considering zero net CO2 emissions associated with biomass
combustion systems, a novel concept has been analysed to capture of CO2 in-situ
with the Ca-looping process while operating the combustor of a dedicated biomass
power plant at sufficiently low temperature. This process is capable of achieving
84% overall CO2 capture rate with an energy penalty of 5.2% when a proper heat
exchanger network is designed with the support of a pinch analysis. The techno-
economic performance of the biomass power plant with in-situ Ca-looping CO2
capture process was compared with that of the alternative biomass-air-fired and
biomass-oxy-fired power plants.
Acknowledgements

First and foremost, I would like to thank my principal supervisor, Prof. Stefano
Brandani without whom this work would not have been possible. This has been a
great opportunity for me, and I appreciate his encouragement and assistance along
the way. I would also like to express my warmest gratitude to my assistant
supervisor, Dr. Hyungwoong Ahn for his support in this dissertation and for teaching
me how to use UniSim software.

I would also thank all those others who contributed to this dissertation by their
comments and suggestions: Prof. Carlos Abanades, Prof. Arturo Macchi, Prof.
Christoph Muller, Prof. Ben Anthony, Dr. Paul Fennell, Dr. Monica Alonso, Dr.
Matteo Romano, Dr. Dennis Lu, Dr. Borja Arias, Dr. Agnieszka Kierzkowska, Dr.
Daniel Friedrich, Dr. Maria-Chiara Ferrari and Dr. Shreenath Krishnamurthy. In
addition thanks to UKCCS Research Center and Carbon Management Canada for
providing me funding to visit INCAR-CSIC, Spain and CanmetENERGY, Canada.
Special thanks for these institutions for hosting me.

I express my indebtedness to the Ministry of Education of the Turkish Government


for providing me the scholarship to conduct this work.

Thanks to everyone in University of Edinburgh, especially to all friends in Carbon


Capture Group and Sanderson Building. I couldn’t have got through without your
support and friendship. I will definitely miss all the times we spent together.

Very importantly thank you to my parents, Nebahat&Mahmut for everything over


the last 30 years. Your support and encouragement has meant everything. Finally,
thank you to my lovely wife Hatice for your kindness, support and patience. I always
feel so lucky to have you in my life.
Table of Contents

LIST OF FIGURES 1

LIST OF TABLES 5

NOMENCLATURE 7

ACRONYMS 11

1 INTRODUCTION 13

1.1 BACKGROUND 13
1.2 CARBON CAPTURE AND STORAGE 14
1.2.1 PRE-COMBUSTION 14
1.2.2 OXY-COMBUSTION 15
1.2.3 CHEMICAL LOOPING COMBUSTION 16
1.2.4 POST-COMBUSTION 17
1.3 TYPES OF CARBON CAPTURE TECHNOLOGY 17
1.3.1 ABSORPTION 18
1.3.2 MEMBRANE SEPARATION 19
1.3.3 CRYOGENIC SEPARATION 20
1.3.4 ADSORPTION 21
1.4 OVERVIEW OF THE CARBON CAPTURE TECHNOLOGIES 21
1.5 CALCIUM LOOPING PROCESS 23
1.5.1 BASICS OF THE CYCLE 23
1.5.2 USE OF SPENT SORBENT FOR CEMENT PRODUCTION 30
1.5.3 PILOT CALCIUM LOOPING PROJECTS 31
1.5.4 ECONOMICS OF THE CALCIUM LOOPING PROCESS 33
1.6 OBJECTIVES OF THE DISSERTATION 33
1.7 ORGANIZATION OF THE DISSERTATION 35
2 PROCESS INTEGRATION OF A CA-LOOPING CARBON
CAPTURE PROCESS IN A CEMENT PLANT 37

2.1 CO2 EMISSIONS FROM CEMENT PRODUCTION 37


2.2 CO2 CAPTURE TECHNOLOGIES FOR THE CEMENT INDUSTRY 38
2.3 PROCESS SIMULATION OF A CEMENT PLANT (BASE CASE) 44
2.4 PROCESS SIMULATION OF A CA-LOOPING PROCESS 50
2.5 PROCESS INTEGRATION OF A CA-LOOPING UNIT WITH A
CEMENT PLANT 55
2.6 RESULTS AND DISCUSSION 59
2.7 CONCLUDING REMARKS 69

3 ALTERNATIVE CO2 CAPTURE TECHNOLOGIES FOR


CEMENT PLANTS 70

3.1 COMBINING CHEMICAL LOOPING COMBUSTION AND


CALCIUM LOOPING PROCESS 70
3.1.1 PRELIMINARY ANALYSIS 72
3.1.2 EXPERIMENTAL DATA FOR THE CLC SORBENT 76
3.1.3 MODIFICATION OF THE RIGOROUS CARBONATOR MODEL
AND PROCESS INTEGRATION 77
3.1.4 RESULTS AND DISCUSSION 81
3.2 INDIRECT CALCINATION PROCESS 87
3.2.1 FUNDAMENTALS OF THE INDIRECT CALCINATION PROCESS 88
3.2.2 PROCESS INTEGRATION OF THE INDIRECT CALCINATION PROCESS 89
3.2.3 AMINE PROCESS AND ITS PROCESS INTEGRATION 91
3.2.4 RESULTS AND DISCUSSION 96
3.3 MEMBRANE SEPARATION PROCESS 103
3.3.1 MODELLING OF THE MEMBRANE PROCESS 103
3.3.2 SELECTION OF AN OPTIMAL FEED GAS STREAM 105
3.3.3 MEMBRANE PROCESS CONFIGURATIONS 108
3.3.4 RESULTS AND DISCUSSION 111
3.3.4.1 SENSITIVITY ANALYSIS OF MEMBRANE PARAMETERS 112
3.4 CONCLUDING REMARKS 115
4 ECONOMIC ANALYSIS 117

4.1 A REVIEW OF EXISTING STUDIES 117


4.2 METHOD OF ECONOMIC ANALYSIS 120
4.3 RESULTS AND DISCUSSION 125
4.3.1 BASE CEMENT PLANT AND CA-LOOPING PROCESS 125
4.3.2 CA-CU LOOPING PROCESS 129
4.3.3 INDIRECT CALCINATION, HYBRID AND STANDALONE
AMINE PROCESSES 132
4.3.4 MEMBRANE PROCESS 136
4.3.5 SENSITIVITY ANALYSIS 138
4.3.5.1 THE EFFECT OF GRID EMISSION FACTOR 143
4.3.5.2 INCLUSION OF FGD AND SCR UNITS IN ALL CONFIGURATIONS 144
4.3.5.3 REPLACEMENT OF THE ASU IN THE CA-LOOPING PROCESS
WITH OXYGEN PURCHASE 146
4.4 CONCLUDING REMARKS 147

5 PROCESS AND COST ANALYSIS OF A BIOMASS POWER PLANT 150

5.1 BACKGROUND 150


5.2 APPLICATION OF CO2 CAPTURE TO BIOMASS-FIRED SYSTEMS 152
5.3 REFERENCE POWER PLANTS 153
5.3.1 BIOMASS-AIR-FIRED POWER PLANT 153
5.3.2 BIOMASS-OXY-FIRED POWER PLANT 157
5.3.3 POWER PLANT AUXILIARIES 157
5.3.4 RESULTS AND DISCUSSION 159
5.4 IN-SITU CA-LOOPING POWER PLANT 160
5.4.1 COMBUSTOR-CARBONATOR MODEL 160
5.4.2 ASSUMPTIONS TO BUILD MASS AND ENERGY BALANCES 161
5.4.3 HEAT RECOVERY STEAM CYCLE 162
5.4.4 RESULTS AND DISCUSSION 168
5.5 ECONOMIC ANALYSIS 170
5.6 CONCLUDING REMARKS 178
6 CONCLUSIONS AND RECOMMENDATIONS FOR

FUTURE WORK 180

6.1 CONCLUSIONS 180


6.1.1 CA-LOOPING PROCESS FOR CEMENT PLANTS 180
6.1.2 CA-CU LOOPING PROCESS FOR CEMENT PLANTS 181
6.1.3 ALTERNATIVE CARBON CAPTURE TECHNOLOGIES FOR
CEMENT PLANTS 181
6.1.4 ECONOMIC ANALYSIS OF THE CARBON CAPTURE
TECHNOLOGIES APPLIED TO CEMENT PLANTS 182
6.1.5 DEDICATED BIOMASS POWER PLANT WITH IN-SITU
CA-LOOPING PROCESS 182
6.2 RECOMMENDATIONS FOR FUTURE WORK 183

REFERENCES 185

APPENDIX A 201

APPENDIX B 211

APPENDIX C 221

APPENDIX D 223

APPENDIX E 227
List of Figures

Figure 1.1 The primary routes for CO2 capture (IPCC, 2005)............................. 15
Figure 1.2 Process schematic of the amine process (CO2CRC, 2014) ................ 18
Figure 1.3 Schematic representation of the separation through a membrane
(Bocciardo, 2014) ............................................................................... 20
Figure 1.4 Cost reduction benefit vs. time to commercialization for innovative
CO2 capture technologies (Figueroa et al., 2008)............................... 22
Figure 1.5 The simplified schematic diagram of the Ca-looping process ........... 24
Figure 1.6 The equilibrium vapor pressure of CO2 over CaO (Garcia-Labiano et
al., 2002) ............................................................................................. 25
Figure 1.7 A cycle of carbonation and calcination observed by a TGA (Ozcan,
2010). The test of absorption capacity was conducted isothermally at
750°C under atmospheric pressure ..................................................... 27
Figure 1.8 CO2 uptake capacities of CaO samples manufactured from the given
precursors at 1000°C in N2 for 1 hr (Ozcan et al., 2011) ................... 28
Figure 2.1 Schematic diagram of a cement plant without a CO2 capture unit
(Base Case) (IEA, 2008). Abbreviations: R/M, Raw Mill; B/F, Bag
Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner .......... 45
Figure 2.2 Clinker phase diagram (Taylor, 1990)................................................ 48
Figure 2.3 Variations of (a) gas and solid temperatures and (b) CO2 concentration
along the cement plant ........................................................................ 55
Figure 2.4 Schematic diagram of the proposed process integration of a cement
plant with a Ca-looping unit. Abbreviations: R/M, Raw Mill; B/F, Bag
Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; Carb,
Carbonator; Calc, Calciner; ASU, Air Separation Unit; CO2 Comp,
CO2 Compression ............................................................................... 58
Figure 2.5 Corresponding FR/FCO2 in range of F0/FCO2 ratio to reach 90% capture
efficiency in the carbonator of proposed configuration (For the
rigorous model, the carbonator temperature and pressure drop along
the column were set as 650°C and 0.1 bar, respectively while
superficial velocity, u0, was estimated to be 6 m/s. The sulfidation
level is shown at each F0/FCO2) ........................................................... 60
Figure 2.6 (a) Energy consumption per unit clinker with respect to fuel and
power and (b) net energy consumption per unit clinker considering
heat recovery for power generation and CO2 compression ................ 62

1
Figure 2.7 Variation of power generation by heat recovery and power
consumption with F0/FCO2 ratio .......................................................... 64
Figure 2.8 Variation of CO2 recovery based on CO2 emission at base case and
incremental energy consumption with F0/FCO2 ratio .......................... 66
Figure 2.9 Variation of carbonation efficiency and incremental energy
consumption per CO2 avoided with F0/FCO2 ratio at 90% CO2
avoidance ............................................................................................ 68
Figure 3.1 Schematic diagram of the conventional CLC process ........................ 71
Figure 3.2 Schematic diagram of the Ca-Cu looping process (Manovic and
Anthony, 2011d) ................................................................................. 73
Figure 3.3 Schematic diagram of the process integration of a cement plant with
an integrated Ca-looping/CLC unit. Abbreviations: R/M, Raw Mill;
B/F, Bag Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-
calciner; Carb, Carbonator; Cal, Calciner (Fuel Reactor); A/R, Air
Reactor; CO2 Comp, CO2 Compression ............................................. 80
Figure 3.4 Experimental data for the CLC sorbent (74.0 wt% CaO, 7.5 wt% CuO
balanced with Al2O3) and natural limestone used in the rigorous
carbonator model (Ozcan et al., 2013) ............................................... 81
Figure 3.5 Corresponding FR/FCO2 values estimated by the simple carbonator
model for the given range of F0/FCO2 to reach 90% capture efficiency
in the carbonator. The results are shown for two different sorbents,
CLC sorbent and natural limestone .................................................... 84
Figure 3.6 The sensitivities of the thermal energy requirement in the capture
plant (MWth), net power capacity (MWe) and incremental energy
consumption (GJth/ton CO2 avoided) against the change in sorbent
performance for the Ca-Cu looping (II) case ..................................... 86
Figure 3.7 Schematic diagram of the indirect calcination process (Rodriguez et
al., 2011b) ........................................................................................... 88
Figure 3.8 Schematic diagram of the process integration of a cement plant with
an indirect calcination process. Abbreviations: R/M, Raw Mill; B/F,
Bag Filter; F/D, Fuel Drying; PHE, Preheater; Cal, CFB Calciner;
Comb, Combustor; CO2 Comp, CO2 Compression ............................ 90
Figure 3.9 Schematic diagram of the amine process (Ahn et al., 2013) .............. 92
Figure 3.10 Process schematic of the CHP plant (IEA, 2008)............................... 93
Figure 3.11 Schematic diagram of the process integration of a cement plant with
(a) hybrid configuration; (b) standalone amine process. Abbreviations:
R/M, Raw Mill; B/F, Bag Filter; F/D, Fuel Drying; PHE, Preheater;

2
Pre-C, Pre-calciner; Cal, CFB Calciner; CHP, Combined heat and
power plant; SCR, Selective Catalytic Reduction Unit; FGD, Flue Gas
Desulphurization Unit; Amine, Amine Scrubbing Process; CO2 Comp,
CO2 Compression ............................................................................... 94
Figure 3.12 Stream properties of the main flows of indirect calcination process
integrated into the base cement plant (Figure 3-8) ............................. 97
Figure 3.13 Variation of solid flux and solid circulation rate in the combustor at
different temperatures ......................................................................... 98
Figure 3.14 Required calcination time at different temperatures and calcination
atmospheres to achieve complete calcination................................... 100
Figure 3.15 Schematic diagram of a cement plant with a membrane-based CO2
capture unit (The end-of-pipe option). Abbreviations: R/M, Raw Mill;
B/F, Bag Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-
calciner; SCR, Selective Catalytic Reduction Unit; FGD; Flue Gas
Desulphurization Unit; CO2 Comp, CO2 Compression.................... 106
Figure 3.16 Schematic diagram of a cement plant with a membrane-based CO2
capture unit (An alternative integration option). Abbreviations: R/M,
Raw Mill; B/F, Bag Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C,
Pre-calciner; SCR, Selective Catalytic Reduction Unit; FGD; Flue Gas
Desulphurization Unit; CO2 Comp, CO2 Compression; HE, Heat
Exchanger ......................................................................................... 107
Figure 3.17 Schematic diagrams of the proposed membrane process
configurations; (a) Conf. 1, (b) Conf. 2, (c) Conf. 3, (d) Conf. 4 ..... 110
Figure 3.18 The impact of sweep/retentate ratio (S/R) on membrane area and
incremental energy consumption...................................................... 113
Figure 3.19 The effects of (a) the CO2 permeance on membrane area; (b) CO2/N2
selectivity on membrane area and energy consumption ................... 114
Figure 4.1 LCOC estimates for the base cement plant, three Ca-looping options
at different F0/FCO2 (see Figure 2-7) and ‘oxy-calciner only’ case. The
error bar reflects the increase in variable cost if there is no surplus
power revenue .................................................................................. 126
Figure 4.2 The cost estimates for the Ca-looping and ‘oxy-calciner only’
processes. For the Ca-looping process, the CO2 avoidance rate is
either fixed at ‘90%’ or ‘>90%’, reflecting the results given in Chapter
2 ........................................................................................................ 128
Figure 4.3 The LCOC estimates for the Ca-looping process using methane as a
fuel (cases A and B) and the Ca-Cu looping process (cases C and D).
Cases A and C refer to the results at 0.02 F0/FCO2 while cases B and D
refer to those at 0.15 F0/FCO2. The error bars reflect the increase in the
variable cost if there is no surplus power revenue ........................... 130

3
Figure 4.4 The cost of CO2 avoided estimates for the Ca-looping process using
methane (A and B) and the Ca-Cu looping process (C and D) ........ 132
Figure 4.5 The LCOC estimates for the standalone indirect calcination, hybrid
configuration and standalone amine process. The error bar reflects the
increase in variable cost if there is no surplus power revenue ......... 135
Figure 4.6 The cost of CO2 avoided estimates for the indirect calcination, hybrid
configuration and standalone amine process. The error bar reflects the
case where there is no revenue for surplus power generation .......... 136
Figure 4.7 The LCOC estimates for Options 1 and 2 at different S/R ratios. The
S/R ratios are shown in the parentheses ........................................... 137
Figure 4.8 The cost of CO2 avoided estimates for the membrane process at two
different feed gas locations ............................................................... 138
Figure 4.9 Sensitivities of (a) TCR and (b) O&M ............................................. 139
Figure 4.10 Sensitivities of (a) fuel cost and (b) power cost ............................... 140
Figure 4.11 Sensitivities of (a) discount rate and (b) scale factor ....................... 142
Figure 4.12 Sensitivity of ETS cost ..................................................................... 143
Figure 4.13 Variation in the cost of CO2 avoided estimate of the Ca-looping option
when the ASU existing in the system is replaced by oxygen purchase
.......................................................................................................... 146
Figure 5.1 Basic schematic diagram of the in-situ Ca-looping plant (Abanades et
al., 2011a) ......................................................................................... 153
Figure 5.2 Simplified schematic diagrams of the (a) biomass-air-fired power
plant (DOE, 2003) and (b) integrated subcritical steam cycle
(IEAGHG, 2009). (HP: High Pressure; IP: Intermediate Pressure; LP:
Low Pressure; FWH: Feedwater Heater) ......................................... 155
Figure 5.3 Schematic diagram of the cryogenic CO2 purification and compression
unit (Xu et al., 2012). (HE: Heat exchanger) .................................. 158
Figure 5.4 Detailed presentation of the heat recovery from the in-situ Ca-looping
plant and its integration with the reference steam cycle shown in
Figure 5-2(b). (The flow rate of steam in the steam cycle is 400 t/h).
.......................................................................................................... 163
Figure 5.5 The grand composite curve prepared according to the specifications
given in Table 5-3............................................................................. 168
Figure 5.6 Sensitivity analysis on cost parameters for the a) biomass-air-fired
plant; b) biomass-oxy-fired plant; c) in-situ Ca-looping plant. (TCR:
Total Capital Requirement; ETS: European Trading Scheme CO2 cost;
GC: Green Certificates) .................................................................... 177

4
List of Tables

Table 2.1 Summary of process configuration models on integration of Ca-


looping process with cement plants .................................................... 41
Table 2.2 Chemical reactions and their standard enthalpies considered in this
cement plant simulation (Taylor, 1990) ............................................. 46
Table 2.3 Composition of the raw meal fed to the raw mill (Taylor, 1990)....... 46
Table 2.4 Comparison of Bogue equation approximation and the simulation
results.................................................................................................. 47
Table 2.5 Composition of the fuels fed to the pre-calciner, kiln and calciner
(* Data from IEA, 2008) .................................................................... 49
Table 2.6 Mass balance of the base case simulation [kg/s] ................................ 50
Table 2.7 Heat balance of the base case simulation [GJth/hr]............................. 51
Table 2.8 Detailed constituents of CO2 avoided energy consumption considering
heat recovery [# Unit: GJth/ton CO2 avoided] .................................... 67
Table 3.1 Chemical reactions defined in the process simulations ...................... 74
Table 3.2 The fitting parameters for the CLC sorbent and limestone ................ 82
Table 3.3 Simulation outputs of the proposed schemes in Section 3.1, including
those for the base cement plant .......................................................... 83
Table 3.4 Main modelling parameters for the amine process............................. 92
Table 3.5 Comparison of the standalone indirect calcination and amine
processes, and the hybrid configuration ........................................... 102
Table 3.6 Composition of the flue gas stream after the bag filters (end-of-pipe)
and the preheater............................................................................... 108
Table 3.7 Comparison of various membrane configurations at 90% overall CO2
avoidance rate ................................................................................... 112
Table 3.8 Minimum incremental energy consumption estimates for the carbon
capture processes evaluated in Chapters 2 and 3.............................. 120
Table 4.1 The main financial assumptions (IEA, 2008) ................................... 122
Table 4.2 Reference capital cost data ............................................................... 123
Table 4.3 Reference variable cost data ............................................................. 125

5
Table 4.4 The composition of the TCRs for the Ca-looping and ‘oxy-calciner
only’ processes presented in Figure 4-1 ........................................... 126
Table 4.5 The composition of the TCRs for the Ca-looping process using
methane (cases A and B) and the Ca-Cu looping process (cases C and
D) presented in Figure 4-3 ................................................................ 130
Table 4.6 The composition of the TCRs for the standalone indirect calcination,
hybrid configuration and standalone amine process ........................ 134
Table 4.7 The composition of the TCRs for the membrane process at two
different feed gas locations ............................................................... 137
Table 4.8 The cost of CO2 avoided estimates using the grid emission factor .. 144
Table 4.9 Variation in the cost of CO2 avoided estimates when the capital costs
of FGD and SCR units are included in all configurations ................ 145
Table 4.10 The summary of cost and process efficiency estimates for the capture
processes evaluated in Chapters 2 and 3. The values in the parentheses
reflect the case where there is no surplus power revenue ................ 149
Table 5.1 Design specifications and performance summaries of the dedicated
biomass-air-fired and biomass-oxy-fired power plants .................... 156
Table 5.2 The stream properties and compositions for the in-situ Ca-looping
plant presented in Figure 5-4 ............................................................ 164
Table 5.3 Pinch analysis specifications for the in-situ Ca-looping plant ......... 167
Table 5.4 Heat balances of the proposed dedicated biomass power plants ...... 169
Table 5.5 The in-situ Ca-looping plant process specifications and performance
summary ........................................................................................... 170
Table 5.6 The main specifications used for the economic analysis.................. 171
Table 5.7 Comparison of the dedicated biomass-fired power plants in terms of
costs of electricity (COE) and CO2 avoided ..................................... 173

6
Nomenclature

a Decay constant of solid concentration in the lean region (m−1)

At Cross sectional area of the reactor (m2)

b Decay constant of contacting efficiency in the lean region (m−1)

CCO2 Actual CO2 concentration in the CFB riser (mol/m3)

CCO2* Equivalent CO2 concentration in the CFB riser (mol/m3)

CCO2,d CO2 concentration in the dense region exit (mol/m3)

CCO2,eq CO2 concentration allowed by chemical equilibrium (mol/m3)

CCO2,in CO2 concentration in the reactor inlet (mol/m3)

CCO2,out CO2 concentration in the reactor exit (mol/m3)

dp Particle diameter (m)

Eac Activation energy of calcination (kJ/mol)

ECO2 Fraction of CO2 captured to total CO2 entering the carbonator

fcarb Extend of carbonation of the particles (Xave/Xmax,ave)

ft Fraction of sorbent particles with a residence time t

F Molar flow rate (kmol/s)

Fash Molar flow rate of ash entering the Ca-looping system (kmol/s)

F0 Make-up (CaCO3) flow rate (kmol/s)

FCO2 Molar flow rate of CO2 in a flue gas stream entering the carbonator (kmol/s)

FR Circulated sorbent (CaO) flow rate (kmol/s)

FS Molar flow rate of sulfur entering the Ca-looping system (kmol/s)

g Acceleration of gravity (m/s2)

Hd Height of the dense region (m)

Hl Height of the lean region (m)

7
Ht Total height of the carbonator (m)

kc Kinetic constant for the calcination reaction (m2kmol−1s−1)

kc0 Pre-exponential factor (m3kmol−1s−1)

kr Kinetic constant of the carbonation reaction (m3/mol/s)

kri Kinetic constant of the carbonation reaction (s−1): krρs,a/Ms,a

kri,ave Average kinetic constant of the population of potentially active Ca-based


solids (s−1)

ks Rate constant for the carbonation reaction at the surface of CaO (m4/mol/s)

Kcw Core-wall mass transfer coefficient (s−1)

Ms Molar mass of total solids (kg/kmol)

Ms,a Molar mass of potentially active Ca-based solids (kg/kmol)

ns,a Moles of potentially active Ca-based solids (kmol)

N Number of carbonation/calcination cycles

p Pressure (Pa)

PCO2,eq Equilibrium CO2 partial pressure (Pa)

rN Mass fraction of particles after N cycles of carbonation-calcination

s Flow rate per unit of width (membrane) (kmol/ms)

SN Specific surface area available after N cycles of carbonation-calcination


(m2/m3)

t Time (s)

tc* Time required to achieve full calcination (s)

tlim Time required for a particle to reach its maximum carbonation (s)

T Temperature (K)

TC Condenser temperature (C)

TH Superheated steam temperature (C)

u0 Superficial velocity of a gas (m/s)

8
v Flow rate per unit of length (membrane) (kmol/ms)

Vg Volume of gas phase (m3)

Vs Volume of solid phase (m3)

Vg,in Volumetric flow rate of inlet gas stream (m3/s)

Vg,out Volumetric flow rate of outlet gas stream (m3/s)

Ws Solid inventory in the carbonator (kg)

x Molar fraction

xash Molar fraction of ash

xCaSO4 Molar fraction of CaSO4

X Carbonation degree of a lime particle

Xave Average carbonation level

Xmax,ave Maximum average carbonation degree of the sorbent after N cycles of


carbonation-calcination

Xmax,N Maximum carbonation conversion rate after N cycles of carbonation-


calcination

Greek letters

ε *s Asymptotic solid volumetric fraction

εs,c Volume fraction of solids in the lower dense region leaner core zone

εs,d Volume fraction of solids in the lower dense region

εs,e Volume fraction of solids at the carbonator exit

εs,l Volume fraction of solids in the lean region

εs,w Volume fraction of solids in the lower dense region denser wall zone

ρg Density of gases (kg/m3)

9
ρs Density of total solids (kg/m3)

ρs,a Density of potentially active Ca-based solids (kg/m3)

ΔXCaSO4 Average fraction of sorbent sulfated at each cycle

μ Viscosity (kg/m/s)

η Net electrical efficiency (kWe/kWth)

ηl The contact efficiency of the lean region

ηsd The contact efficiency of the dense region

ηT Steam turbine efficiency

δ The core volumetric fraction of the dense region

ξ Volume ratio between the potentially active solids and the total solids

τ Average residence time of solids particles (s)

π Permeance (GPU)

10
Acronyms
ASU Air separation unit

BECCS Bio-energy with CO2 capture and storage

BFB Bubbling fluidized bed

Ca-looping Calcium looping process

CCS Carbon capture and storage

CF Capacity factor

CFB Circulating fluidized bed

CHP Combined heat and power

CLC Chemical looping combustion

COC Cost of cement

COE Cost of electricity

COM Component object model

COP Coefficient of performance

ESP Electrostatic precipitator

ETS Emission trading system

FC Fuel cost

FCF Fixed charge factor

FGD Flue gas desulphurization

FWH Feedwater heater

GC Green certificate

GCC Grand composite curve

GPU Gas permeation unit

IGCC Integrated gasification combined cycle

LCOC Levelised cost of cement

LCOE Levelised cost of electricity

11
LHV Lower heating value

M&S Marshall&Swift cost index

MEA Monoethanolamine

MER Minimum energy requirement

MOFS Metal-organic frameworks

O&M Operating & maintenance

OTM Oxygen transport membrane

PC Pulverised coal

PSA Pressure swing adsorption

ROC Renewable obligation certificates

SCR Selective catalytic reduction

TCR Total capital requirements

TGA Thermal gravimetric analysis

TSA Temperature swing adsorption

VOM Variable operating and maintenance

12
Chapter 1 Introduction
1.1 Background

Global CO2 emissions hit a new record of 34.5 Gt in 2012 (Olivier et al., 2013)
and are predicted to increase further, reaching 40.2 Gt by 2030 (IEA, 2009a). The
National Oceanic and Atmospheric Administration (NOAA)/Earth System Research
Laboratory reported that the global CO2 concentration reached 400 ppm in 2013,
representing an increase of 24% from 1958 (Dlugokencky and Tans, 2014).
According to the latest Intergovernmental Panel on Climate Change report (IPCC,
2014), it is more certain than ever that CO2 emission from fossil fuel burning is the
major contributor to changing the earth’s climate. Unless alternative energy systems
are developed and rapidly deployed, it is expected that fossil fuels will remain as the
major contributor for energy production (IEA, 2012) while world energy demand is
envisioned to increase by 40% from 2007 to 2030 (IEA, 2009b).

The increasing demand in electricity will trigger CO2 emissions dramatically if


applicable technologies to reduce such emissions are not developed. For the sake of
restraining the severe effects of climate change and limiting the increase in the
atmospheric temperature to 2°C (EG Science, 2008), substantial reductions of global
CO2 emissions by at least 50% are required by 2050 (IPCC, 2007). The UK aims for
at least 80% reduction in its CO2 emissions relative to 1990 levels by 2050 as stated
in the Climate Change Act (UKP, 2008). Moreover, the Kyoto Protocol ratified in
1997 (Kyoto, 1997) and the Bali Declaration issued at the United Nations Climate
Change Conference in 2007 (UNFCC, 2007) also emphasize the requirement for
reductions in greenhouse gases, especially for CO2 which accounts for 77% of the
total anthropogenic emissions (IEA, 2009c) originating from the combustion of fossil
fuels and from transport/households.

Being one of the primary environmental targets, the reduction of net CO2
emissions to the atmosphere can be achieved by (IPCC, 2005): (i) improving the
efficiency of energy conversion and/or utilization to reduce energy consumption, (ii)
inducing the use of renewable and less carbon intensive fuels, (iii) enhancing the
biological absorption capacity in forests and soils and (iv) capturing and storing CO2.

13
1.2 Carbon Capture and Storage

Carbon capture and storage (CCS) is a promising and an emerging way of


reducing CO2 emissions from the main contributors: fossil-fuelled power stations and
industrial processes such as cement, iron and steel production plants (DOE, 1999).
The IEA (2010) estimates that up to 19% reductions in CO2 emissions by 2050 can
be achieved when CCS is applied to these main emission sources. In the framework
of CCS, CO2 emissions from a process are captured and the high concentration CO2
product is separated. Thus, part of the CO2 (depending on the efficiency of a carbon
capture process) is not emitted to the atmosphere.

The captured CO2 can be injected into geological formations (Torp and Gale,
2004) but is also used for enhanced oil recovery (Jaramillo et al., 2009). The
technology needed to deploy CCS already exists in the natural gas, oil and chemical
industries, where CO2 and other gases such as H2 and O2 have been separated from
different gas mixtures (Rodriguez et al., 2010). It has been proven by an economic
analysis given in Rao and Rubin (2002) that CO2 capture accounts for 75-85% of the
overall cost associated with CCS. Therefore, it is important to develop efficient and
cost-effective carbon capture technologies, which are usually split into three general
routes: pre-combustion, oxy-combustion and post-combustion as presented in Figure
1-1. In addition, chemical looping combustion (CLC) is a new generation
combustion technology with inherent CO2 capture. These technologies are briefly
detailed in the sections that follow.

1.2.1 Pre-combustion

The main objective of this technology is to convert a fuel to carbon-free H2 prior


to combustion. It involves three steps. First, a hydrocarbon containing fuel such as
coal or biomass reacts with steam or air and/or oxygen to produce syngas or fuel gas.
Second, CO generated in the first step catalytically reacts with steam to generate CO2
and more H2 through the water gas shift reaction. Finally the product gas containing
mainly CO2/H2 proceeds to a carbon capture process where CO2 is separated from
H2.

14
Figure 1-1 The primary routes for CO2 capture (IPCC, 2005).

The potential carbon capture technologies for this purpose include adsorption,
absorption and membrane separation. The resulting H2-rich product can be fired in
boilers, furnaces, heaters or as a fuel in vehicles (Harrison, 2008; Pennline et al.,
2008). This technology is close to commercial reality in the form of the Integrated
Gasification Combined Cycle (IGCC) (CCSGI, 2010) where the use of physical
sorbents, such as Selexol (Kapetaki et al., 2013) is preferable because of the high
temperature and pressure gas streams. Nevertheless, since only CO2 emissions from
fuel combustion can be avoided by pre-combustion technologies, only moderate
carbon capture efficiencies can be achieved when this technology is applied into the
cement industry, where CO2 is generated by two different sources: fuel combustion
and calcination of limestone. In general pre-combustion is only applicable for new
plants rather than existing ones, which is the opposite of post-combustion for which
retrofit integration is often possible.

1.2.2 Oxy-combustion

In this technology, almost pure oxygen (≥95 mol%) is supplied to a combustor


instead of air, so that it is in principle possible to increase the CO2 concentration in
the flue gases to more than 90 mol% after condensation of water vapour compared
with 10-15 mol% for air-based combustion systems (Jared et al., 2010). To maintain
the same boiler flame temperature as in air-combustion, part of the exit gas stream is

15
generally recycled to lower the concentration of the inlet oxygen. The type of fuel,
the excess quantity of oxygen and the potential air-leakages are the main elements
determining the final CO2 purity. Consequently, a CO2 purification stage could be
needed to increase the purity further to meet storage specifications (≥95 mol%). The
currently available oxy-combustion technologies are based on cryogenic air
separation unit (ASU) for oxygen production, which is an expensive and energy-
intensive process. Promising progress on the development of membrane-based
systems for oxygen production to reduce power consumption by 70% to 80%
compared to a cryogenic system has been reported (Stadler et al., 2011). The
Babcock & Wilcox Company has already demonstrated oxy-combustion of coal on a
30 MWth boiler (McCauley et al., 2009).

1.2.3 Chemical Looping Combustion

CLC is a new generation energy technology allowing inherent production of pure


CO2 after condensation of steam (Lyon and Cole, 2000; Ishida and Jin, 2004). In
contrast to the conventional combustion process, oxygen for combustion is provided
by the reduction of metal oxides in a fuel reactor as presented in the following
reaction. Some well-known oxygen carriers are the oxides of Ni, Cu, Fe and Mn
supported by inert materials such as Al2O3, MgO and TiO2 (Hossain and de Lasa,
2008). The exit gas stream from the fuel reactor contains CO2 and H2O.

(2𝑛 + 𝑚)𝑀𝑒𝑥 𝑂𝑦 + 𝐶𝑛 𝐻2𝑚 → (2𝑛 + 𝑚)𝑀𝑒𝑥 𝑂𝑦−1 + 𝑚𝐻2 𝑂 + 𝑛𝐶𝑂2 (Rn. 1.1)

H2O can then be condensed to obtain pure CO2. The reduced metal oxide from the
fuel reactor is then re-oxidized with air in an air reactor.

𝑀𝑒𝑥 𝑂𝑦−1 + 1�2 𝑂2 → 𝑀𝑒𝑥 𝑂𝑦 (Rn. 1.2)

The total amount of heat generated by reduction and oxidation in the two reactors
is equal to that of combustion of a fuel. The main advantages of the CLC process
compared to the oxy-combustion technology are the absences of an ASU and the
excess oxygen for combustion that reduces the CO2 purity. A considerable number of
studies have been conducted to improve the stability and performance of the metal

16
oxides in use (Leion et al., 2009; Adanez et al., 2012). Mature circulating fluidized
bed (CFB) technology is well-suited for the application of the CLC process.

1.2.4 Post-combustion

Post-combustion refers to the separation of CO2 from any flue gas streams
containing CO2 mixed with other gases. It can be adapted to any industrial process
emitting large quantities of CO2 as an end-of-pipe solution without needing excessive
modifications to the actual process scheme. There is a vast potential for retrofitting
post-combustion technologies to large anthropogenic CO2 emission sources in the
UK where nearly a third of electricity is generated from coal (Mellows-Facer, 2010).
In addition, the IPCC (2005) indicated that approximately 8000 large stationary CO2
sources around the world releasing 14 Gt CO2 per year could potentially be
retrofitted with CCS.

The most common types of post-combustion technologies are summarized in the


next section. However, it should be first highlighted that the leading post-combustion
carbon capture technology should provide high capture efficiencies with less
expense, meaning lower energy and cost requirements. Therefore, one of the
milestones in the development of these technologies is the evaluation of their techno-
economic performance. For this purpose, the transfer of knowledge from
experimental studies would enable the development of reactor models, which can
then be employed in fully integrated process flowsheets.

1.3 Types of Carbon Capture Technology

The most common methods for the separation of CO2 from gas streams include:

1) Absorption

2) Membrane Separation

3) Cryogenic Separation

4) Adsorption

17
Figure 1-2 Process schematic of the amine process (CO2CRC, 2014).

1.3.1 Absorption

This technology relies on the principle that CO2 in the feed gas is transferred into
the liquid phase by selective absorption in a solvent. The solvent rich in CO2 can
then be regenerated by the application of heat while pure CO2 is released. The
absorption process typically works at low temperatures, between 40°C (absorption)
and 120°C (desorption). The CO2 separation can be achieved either physically or
chemically. Physical absorption relies on the selective solubility of different gases
while the solubility of the absorbent increases at high partial pressures. In chemical
absorption (or reactive absorption), however, CO2 reacts with the absorbent and
better absorption performance can be achieved at low partial pressures.

Amine process, based on removal of acidic gases from gas streams by chemical
absorption with aqueous amine solutions, is the most mature technology for post-
combustion CO2 capture applications from low pressure and low CO2 concentration
flue gases. It has been used for over 80 years (Bandyopadhyay, 2011) and several
techno-economic studies are available for this system (Finkenrath, 2011; EPRI,
2009). The well-known and most widely used amine is monoethanolamine (MEA).
The base schematic diagram for the amine process is shown in Figure 1-2 where the
CO2 in flue gases reacts chemically with the solvent in an absorber, and the solvent is

18
regenerated in a desorber (stripper). The clean flue gases from the absorber, a vertical
and packed column, generally go through a water wash tower to remove any MEA
slip. The CO2 rich solvent from the absorber is fed into the top of the stripper column
after being heated up in a heat exchanger. The chemically bound CO2 is separated
from the solvent in the stripper while heat is provided in a reboiler.

The level of NOx and SOx concentrations in the feed gas stream is crucial for the
amine process. These components react with amine to produce amine salts which
cannot be dissociated in the stripper. The spent solvent needs to be replaced with
fresh solvent which increases the variable cost. Therefore, flue gas desulphurization
(FGD) and selective catalytic reduction (SCR) units need to be included in the
process to limit SOx and NOx emissions. Applying this technology within a power
plant, the heat requirement in the stripper is usually provided by the steam dilution
from steam cycle of the power plant. In Ahn et al. (2013), ten different advanced
amine process configurations were systematically evaluated for the purpose of CO2
capture from a power plant, and the heat duty in the stripper was reduced to as low as
2.22 MJth/kg CO2 from 3.52 MJth/kg CO2 for the conventional system by enhancing
heat recovery and increasing solvent working capacity.

1.3.2 Membrane Separation

Gas separation using membranes relies on selective and specific permeation of


different gases. The driving force for the separation can be expressed as the
difference in partial pressure between the two sides (retentate and permeate), and
feed compression and/or vacuum at the permeate side are often adopted. The
schematic representation of the membrane separation principle is presented in Figure
1-3. The permeability represents the ability of the molecules to permeate through the
membrane film and can be written as the product of diffusivity and solubility
according to the solution diffusion theory (Baker, 2002; Baker, 2004). For effective
removal of CO2 from a gas stream, membrane materials should possess a number of
properties including high CO2 permeability and selectivity as well as thermal and
chemical stability.

19
Figure 1-3 Schematic representation of the separation through a membrane (Bocciardo, 2014).

Currently this technology is only applied to the removal of CO2 from natural gas
but numerous membranes have been developed and commercialised for gas
separations since the 1970s (Robeson, 2008). Dense polymeric membranes are the
main candidates for post-combustion carbon capture (Merkel et al., 2010).
Unfortunately it was reported that the selectivity of membranes can possibly be
affected by impurities, i.e. SOx, NOx and H2O in flue gases (Scholes et al., 2011).
Furthermore multiple stages and considerable membrane areas would be required to
achieve high capture efficiencies and CO2 purities (e.g. ~90% capture and over 95
mol% purity). This technology is still not available commercially for CO2 capture
applications in spite of promising results from pilot-scale investigations (NETL,
2012).

1.3.3 Cryogenic Separation

For the separation of CO2 from other gases, the cryogenic separation process uses
low temperatures and high pressures to reach a condition where CO2 is in liquid
phase. The operating conditions in this system are maintained by the use of
compressors and heat exchangers in series. This technology is commercially used for
the purification of gas streams at high CO2 concentrations (>90 mol%). However, the
major disadvantage of cryogenic separation of CO2 is the excessive energy
requirement for the refrigeration step, particularly at very low CO2 concentrations. It
has an advantage in that the separated CO2 by this system is already in liquid phase
which is beneficial for certain transport options.

20
1.3.4 Adsorption

Adsorption is a cyclic process where CO2 in a gas stream is initially adsorbed on


the surface of a solid either physically or chemically and thereby removed from the
gas stream. In physical adsorption, the bonding between the CO2 and the surface is
due to weak van der Waals forces, and low temperatures are favourable whereas
strong chemical bonding in chemical adsorption is favourable at higher temperatures
(>200°C). The main types of adsorption processes involve pressure swing adsorption
(PSA) and temperature swing adsorption (TSA). In PSA, the regeneration is provided
by the difference in pressure whereas it is the temperature which allows regeneration
of the sorbent in TSA. The currently available most effective sorbents are activated
carbons, zeolites and aluminas. Researchers have been working on a large scale
application of this technology in the industrial process, which requires highly
selective and efficient adsorbents (Harlich and Tezel, 2004; Boot-Handford et al.,
2014; Hu et al., 2014). The need for sorbents with elevated performance
characteristics leads to the development of novel sorbents.

1.4 Overview of Carbon Capture Technologies

Figure 1-4 illustrates the potential of CO2 capture technologies for


commercialization and cost reduction benefit. Currently amine-based post
combustion technology is closest to the market. However, the integration of an amine
process into an existing power plant is energetically intensive and economically
prohibitive. The reason for the latter is also related to the necessity to prevent solvent
degradation as a result of reaction with SOx and NOx in flue gases in addition to the
high cost of solvents. Reports suggest that the SO2 concentration in the flue gas
needs to be maintained below 10 ppm before an amine process, revealing the
requirement of over 98% SO2 removal even for the lowest sulphur coals (Rao and
Rubin, 2002). Another potential technology that is close to the market is the oxy-
combustion process. Nevertheless, this process requires an energy intensive ASU to
produce significant amounts of oxygen for combustion. In addition to that, very
stringent safety management along with the control of air-leakages is needed.

21
Figure 1-4 Cost reduction benefit vs. time to commercialization for innovative CO2 capture
technologies (Figueroa et al., 2008).

For some of the next generation technologies, for example, ionic liquids, metal-
organic frameworks (MOFs) and oxygen transport membrane (OTM) boilers, a
considerable number of investigations at pilot scale have been continuing but
significant R&D efforts are still needed before these technologies reach large-scale
applications. Because of such deficiencies associated with most conventional CO2
capture technologies, there is an urgent need to develop more advanced and less
expensive methods by which the above issues can be minimized.

One of such promising technologies for carbon capture from industrial sources is
an absorption process based on the reversible reaction of CO2 on specific metal
oxides at high temperatures. The CaO-based sorbents have attracted the most
attention owing to their high absorption capacity, wide availability and low cost. The
concept of utilizing CaO for CO2 capture was first introduced by DuMotay and
Marechal for enhancing coal gasification in 1869 (Squires, 1967) and named the
calcium looping (Ca-looping) process. The Ca-looping process is applicable to pre-

22
combustion (Albrecht, 2008a; Weimer et al., 2008; Blamey et al. 2010) and post-
combustion (Shimizu et al., 1999; Abanades and Alvarez, 2003; Abanades et al.,
2005; Berstad et al., 2012) CO2 capture applications. It has attracted great interest
owing to the use of a cheap sorbent, a relatively small energy penalty (6-8%)
compared to 8-10% for post combustion amine scrubbing (Florin and Fennell, 2010),
the potential use of large scale CFB systems as a mature technology, the reduction of
ASU power consumption (approximately 1/3 of that for the oxy-combustion)
(Abanades et al., 2005), and the possible use of the purge stream in cement
manufacturing plants which has a potential to improve the economics of the system
and to de-carbonize both industries.

1.5 Calcium Looping Process

1.5.1 Basics of the Cycle

The post-combustion Ca-looping process that was first proposed by Shimizu et al.
(1999) uses state-of-the-art CFB systems due to the huge flue gas flow coming into
contact with a solid stream as presented in Figure 1-5. In this process, CO2
containing flue gases are fed to a reactor, called carbonator, and CO2 is captured at
600-750°C according to the following reaction:

CaO + CO2 ↔ CaCO3 (ΔH923 K = −171 kJ/mol) (Rn. 1.3)

The solids from the carbonator are separated in a cyclone and sent to a calciner
where the product of carbonation, CaCO3 is regenerated and high purity CO2 at 870-
950°C is obtained. The heat requirement in the calciner, which is a relevant fraction
(35-50%) of the total energy entering the entire system (i.e. the power plant and Ca-
looping process together) (Abanades et al., 2007; Rodriguez et al., 2008a), is often
provided by oxy-combustion of a fuel (Romeo et al., 2008; Romeo et al., 2009).
Otherwise, if air is used for combustion, the N2 from the air dilutes the gas stream
from the calciner. The use of transfer mediums, such as hot CaO particles from a
high temperature combustor has also been suggested as a means of providing heat for
the calciner (Abanades et al., 2005).

23
CO2-depleted
Gas
CO2-rich Gas

Carbonator Calciner

Flue Gas Oxygen


Fuel
Make-up

Purge

Figure 1-5 The simplified schematic diagram of the Ca-looping process.

The thermal energy requirement in the calciner depends mainly on the flow rate of
make-up sorbent and solid circulation rates between the reactors, which are directly
linked to overall capture efficiency. Despite the fact that additional heat is required in
the calciner, one of the advantages of this system is the possibility of recovering
surplus energy from hot gas and solid streams leaving the reactors as well as the
exothermic carbonation reaction. The recovered heat can then be used to drive a
steam cycle to generate electricity, reducing the energy penalty of the system.

A plot of temperature vs. equilibrium CO2 vapour pressure (Garcia-Labiano et al.,


2002) is given in Figure 1-6. This shows that above the equilibrium line carbonation
is favourable while below the line CaCO3 decomposes into CaO and CO2. The graph
also provides a convenient means for the estimation of the minimum calcination
temperature which depends on CO2 partial pressure inside the calciner. In order to
produce a suitable gas from the calciner for storage, the calciner should operate at
high CO2 partial pressures. According to CaO-CaCO3 equilibrium, high CO2 partial
pressure in the calciner imposes a temperature of more than 900°C if the calciner
operates at atmospheric pressure. For a calciner operating at atmospheric pressure,
calcination starts at around 600°C under pure nitrogen but the starting temperature
delays to 880°C under pure CO2 (Ozcan, 2010).

24
1

0.1

PCO2 (atm)
0.01

0.001

0.0001
550 600 650 700 750 800 850 900
T (°C)

Figure 1-6 The equilibrium vapor pressure of CO2 over CaO (Garcia-Labiano et al., 2002).

Even though there are various experimental studies where the CO2 uptake
performance of a CaO-based sorbent was evaluated under mild calcination
conditions (Salvador et al., 2003; Ryu et al., 2006), the need for CaCO3 calcination in
a CO2 rich atmosphere (>90%) for production of a high purity CO2 stream
significantly affects the carbonation behaviour of the sorbent. Thus, the calciner
temperature is often determined by a compromise between the production of high
CO2 partial pressure and the expense of sorbent degradation. At high temperatures,
calcination kinetics is very fast which allows complete calcination to occur rapidly,
but also high calcination temperatures trigger loss in the sorbent capacity because of
sintering. Sintering is a term referring to changes in the sorbent structure, pore shape,
pore shrinkage and grain growth. Similarly, the carbonator reaction is governed by
the equilibrium between temperature and CO2 partial pressure. The effect of sintering
is more prominent at high temperatures and in severe calcination conditions (high
steam and CO2 partial pressures) (Abanades and Alvarez, 2003; Chrissafis et al.,
2005). Thus, sintering of CaO is more effective in the calciner than the carbonator
because of the operating conditions.

25
A method of reducing CO2 partial pressure in the calciner is the dilution of the
calcination environment with steam. It allows the reduction of the required
calcination temperature for complete calcination and prevents calcination at high
temperatures (Garcia-Labiano et al., 2002). The steam in the CO2-rich gases can then
be condensed and separated from the CO2. Khinast et al. (1996) indicated that the
calcination rate increases exponentially with decreasing CO2 partial pressure at a
constant temperature. Wang et al. (2008) illustrated that the calcination conversion of
72% at 100% CO2 atmosphere increases to 98% with 60% steam dilution at a bed
temperature of 920°C and average residence time of 40 mins. It should be noted that
the selection of a steam source to dilute calcination atmosphere can be challenging. If
it is need to be taken from the steam cycle of a power plant, the steam dilution results
in a decrease in the output of the power plant and is thus a potential energy penalty
for the capture process. Another solution to minimize the calcination temperature is
the application of a vacuum to reduce pressure in the calciner (Abanades et al.,
2005). It has been shown that the calcination of a sorbent under vacuum conditions at
a lower temperature improves the kinetics of the sorbent by reducing the effect of
sintering (Skadjian et al., 2007). However, the application of a vacuum is expensive
especially at larger scales.

As presented in Figure 1-7 for a cycle of carbonation-calcination reaction (Ozcan,


2010), the carbonation reaction contains two steps. After an initial rapid, kinetically
controlled first stage, the second stage is very slow and diffusion controlled (Bhatia
and Perlmutter, 1983). The calcination stage generally proceeds to completion. The
materials used in the Ca-looping process can be any widely available and low cost
natural sorbents such as limestone and dolomite (Silaban et al., 1996; Grasa et al.,
2008b). A CaO-based sorbent can also be manufactured from synthetic precursors
such as Ca(OH)2, Ca(C2H3O2)2 etc. Grasa et al. (2007) demonstrated that the sorbents
derived from synthetic CaO precursors present similar behaviour to the limestone
under realistic calcination conditions (high temperature and CO2 partial pressure)
even though these sorbents show much better stability and performance under mild
calcination conditions.

26
60
Kinetically- controlled Diffusion-controlled
carbonation stage carbonation stage

Weight (mg) 50

40 Calcination Stage

30
0 10 20 30 40 50 60 70
Time (min)

Figure 1-7 A cycle of carbonation and calcination observed by a TGA (Ozcan, 2010). The test of
absorption capacity was conducted isothermally at 750°C under atmospheric pressure.

One of the greatest challenges in the Ca-looping process is the deterioration of the
CO2 capture capacity when these materials are used over a number of cycles of
carbonation and calcination reactions (Curran et al., 1967; Silaban and Harrison,
1995; Grasa and Abanades, 2006). For continuous processes, it would be preferable
to use a sorbent for countless numbers of cycles. However, the evolution of the
capture capacity of natural sorbents reveals that the capture capacity significantly
decreases during the first 20 cycles but then stabilizes at around 8% even after 500
cycles (Grasa and Abanades, 2006). For the carbonation reaction, a sufficiently large
pore volume needs to be ensured but the available pore volume in the sorbent derived
from limestone decreases through the cycling progress. Figure 1-8 shows the CO2
uptake capacities of CaO samples prepared from different precursors: pure CaO,
Microna 10 limestone, dolomite and calcium acetate (Ozcan et al., 2011). The
samples demonstrate a decrease in their absorption capacities through the cycles
while that of dolomite seems more stable compared to others owing to MgO acting
as a stabilizer. The major reason for lowering residual activity was sintering of
experimentally tested sorbents at high temperatures while the potential challenges
could be extended to include sulphation and attrition in large scale applications of the
Ca-looping process.

27
16

14

12
Absorption Capacity,
mmol CO2/g sorbent 10

4 Pure CaO
Microna 10
2 Dolomite
Calcium Acetate
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Cycle

Figure 1-8 CO2 uptake capacities of CaO samples manufactured from the given precursors at 1000°C
in N2 for 1 hr (Ozcan et al., 2011).

CaO-based sorbents have a very strong affinity to absorb SO2 existing in the flue
gases and generated by coal combustion in the calciner in the form of CaSO4, as
shown in (Rn. 1.4), because of very high Ca/S ratio in the reactors.

CaO + SO2 + ½ O2 → CaSO4 (ΔH = −506 kJ/mol) (Rn. 1.4)

CaSO4 is stable under the conditions of carbonation and calcination, thus, in order to
regenerate CaSO4 back to CaO, either high temperatures (>1250°C) or a reducing
atmosphere is needed. The reaction between CaO and SO2 can be considered as loss
in reactive sites for carbonation in the Ca-looping process. Sun et al. (2007) indicated
that sulphate formation results in pore blockage as it has a much larger volume and
prevents CO2 flow to the centre of a particle. On the other hand, CaO derived from
calcination of natural limestone has been commercialized in CFB power plants for
sulphur control (Srivastava et al., 2001). In such systems, there is no need for an
external FGD unit, which is necessary for pulverised-coal (PC) combustion power
plants, unless the emission regulations are extremely strict. The absence of FGD can
be, therefore, counted as profit for its capital cost. Furthermore, the attrition of the

28
limestone can be problematic in large scale CFB systems where the CaO particles
can break apart by collision with other particles, and resulting fine particles escape
the system with the gas stream (Coppola et al., 2012). In a recent study, pelletisation
of artificial sorbents using calcium aluminate cements as binders has presented both
high reactivity and resistance against the attrition (Manovic and Anthony, 2009a).

Substantial research efforts are currently being undertaken to reduce the decay in
reactivity of CaO or to improve the reactivity of the used sorbent (Lu et al., 2006;
Manovic and Anthony, 2009b). These methods include thermal pre-treatment
(Manovic and Anthony, 2008; Arias et al., 2011), reactivation of spent sorbent by
hydration (Fennell et al., 2007; Sun et al., 2008; Manovic and Anthony, 2011a;
Wang et al., 2012) and doping of natural sorbents with inert materials (Manovic and
Anthony. 2009c; Sun et al., 2012).

Lysikov et al. (2007) investigated the effects of thermal activation on sorbent


performance. It was observed that prolonged exposure of the sorbent derived from
limestone to severe calcination/carbonation conditions is beneficial for the cyclic
CaO conversion. The severely sintered sorbent initially possesses a low CO2 uptake
capacity in the first cycles because of reduced pore volume but the conversion
actually increases later, which is referred to self-reactivation. Recently, it was
indicated that the improvement in the CO2 uptake capacity of the thermally pre-
treated sorbents is not significant after 100 cycles (Ozcan et al., 2011). It was proven
that the reactivity of hydrated limestone is even better than fresh limestone (Manovic
and Anthony, 2007); however, the side effect of hydration was reported to be the
deterioration of the mechanical strength of reactivated sorbents resulting in high rates
of attrition (Manovic et al., 2008).

Another approach for improving the cyclic stability of the sorbent involves the
incorporation of the limestone with a variety of dopants such as Al (Pacciani et al.,
2008), Mg (Albrecht et al., 2008b) or Si (Huang et al., 2010) because of their high
Tammann temperatures in the sorbent. The Tammann temperature is approximately
half of the melting point and is often considered as the point where sintering
becomes rapid. The incorporated inert material is expected to suppress the sintering.

29
The inert supports often do not react with CaO, forming mixed oxides but the reverse
is also possible (Manovic and Anthony, 2010).

Unless highly stable and efficient CaO-based sorbents are derived, some spent
sorbents need to be removed as a purge from the system while fresh sorbent is
introduced. From the viewpoint of process economics, there is a trade-off between
the sorbent cost and performance. In general, the production of synthetic sorbents is
more costly compared to natural sorbents. Thus, the expectation would be
manufacturing synthetic sorbents that possess high and stable CO2 uptake capacities
to minimize the requirement of the sorbent replacement. However, if this is not
achievable, the use of natural sorbents with affordable purge flow rates would be
preferable.

1.5.2 Use of Spent Sorbent for Cement Production

The purge stream from the coal-fired calciner fed with limestone is mainly
composed of CaO with small amounts of CaSO4 and ash. According to Hurst et al.
(2012), a coal-fired power plant integrated with a Ca-looping process could be a
carbon negative if the purge sorbent is used in the cement industry or disposed in the
ocean forming bicarbonate. If the spent sorbent from the Ca-looping process can be
used as a raw material for cement manufacture, the CO2 released by the calcination
of limestone in the cement process can be eliminated. Moreover, the heat
requirement for calcination and the costs of limestone and fuel in the cement plant
can be reduced.

Even though it was conceivable in earlier studies that the purge CaO could be sent
to a kiln to make a clinker, there has been an issue that the deactivated CaO may
deteriorate the clinker quality and the factors which can potentially limit or affect the
re-use of purge CaO have been discussed (Dean et al., 2011a). The study raised
issues that sulphur conveyed by a purge stream can lead to expansion and cracking of
the cement paste upon hydration and also affects the formation of cement phases.
Moreover, the trace elements released from fuel combustion in the calciner as well as
attrition and agglomeration have been identified as other potential issues in using
purge materials. In the latter work of this group, it has been shown experimentally

30
that cement can be successfully produced from the purge by utilizing CaO which has
experienced many cycles of calcination and carbonation (Dean et al., 2011b). In a
more recent work (Telesca et al., 2014), the spent sorbent from a pilot scale Ca-
looping plant, purged after 60 cycles of carbonation and calcination, was employed
in place of limestone in a lab-scale cement manufacturing process. It was concluded
that the burnability of this mixture is better than one including fresh limestone while
the mineralogical composition and hydration behaviour of both samples were found
to be identical.

1.5.3 Pilot Calcium Looping Projects

In addition to outstanding efforts in lab-scale investigations (Alonso et al., 2010;


Rodriguez et al., 2011a), several different projects around the world have been
initiated to scale up the Ca-looping technology, including a 1.7 MWth pilot plant
which has been in operation to test the concept in La Pereda, Spain funded under
“CaOling” project since January 2012 (Arias et al., 2013). The pilot plant treats
1/150 of the flue gases emitted from La Pereda CFB power plant and includes two
interconnected CFB reactors: a carbonator and an oxy-fired calciner, which are 15 m
in height. It has been operated more than 380 hours in steady state mode and
demonstrated to achieve CO2 capture efficiencies over 90% with the supply of
sufficient CaO. Furthermore, promising results have also been reported from a 1
MWth pilot plant in Darmstadt (Plötz et al., 2012). In this plant, the capture
efficiencies over 80% have been achieved in the carbonator (overall >90%) for
several hours of operation, and the limitations of reactions kinetics and chemical
equilibrium on CO2 capture efficiency have been demonstrated by changing process
parameters. The heat requirement in the calciner is provided by combustion of either
propane or pulverised hard coal. The height of the CFB carbonator was set to 8.7 m
while that for the calciner is 11.3 m, where a maximum flow of 150 kg/h can be
introduced.

Initial experience on 10 kWth facility led IFK Stuttgart to scale up their Ca-
looping plant to 200 kWth (Dieter et al., 2004). The main purpose of the large scale
Ca-looping plot is to investigate different fluidization regimes and concepts for solid

31
circulations. Three interconnected fluidized bed reactors were built, which provides
flexibility to operate the carbonator in two modes: either a fast fluidized bed or a
turbulent fluidized bed. While the calciner is 10 m in height, the carbonator is either
10 m (the fast fluidized) or 6 m (the turbulent fluidized). Even though it was
concluded that the turbulent reactor offers greater flexibility for the variations in flue
gas load compared to the fast fluidized bed, the latter was reported to allow more flue
gas load and has been preferred at equal active space times because of better
performance. Moreover, a great attention has been given to sorbent attrition which
was defined as one major hurdle during the operation. A minimum 3 wt%/h solid
inventory lost was observed regarding to the attrition.

Canmet Energy, Canada operates 75 kWth dual fluidized bed system (Lu et al.,
2008). The carbonator was constructed as a BFB, and the calciner is an oxy-fired
CFB. The calciner can be operated under three different modes: oxy-combustion of
biomass or coal and electrical heating. With continuous operations more than 50 h,
promising CO2 capture efficiencies have been achieved. Attrition of sorbent was also
defined as the major problem in this plant while the attrition rate was found to be
~50% for Havelock limestone particles, in the range 0.4-0.8 mm, which eventually
became less than 0.1 mm.

A 1.9 MWth Ca-looping pilot designed by ITRI, Taiwan is currently being erected
(Chang et al., 2013). The design of this pilot was based on experimental results
obtained by a 3 kWth bench-scale system. The carbonator was built as a BFB with a
diameter of 3.3 m and height of 4.2 m, and operates at 650°C with gas residence time
of 8 – 10 s. The 36 water-cooled double steel jackets were installed to remove excess
heat from the carbonator because of the exothermic carbonation reaction and the high
temperature solid stream from the calciner. The calciner was designed as a rotary
kiln, as existing in the cement manufacturing process, and 5 m in length. It was
facilitated with an oxy-combustion system including flue gas recirculation. Even
though the system is able to run smoothly at this stage, the evaluation of process
parameters is still under investigation.

32
1.5.4 Economics of the Calcium Looping Process

Abanades et al. (2007) reported the cost structure for a post-combustion Ca-
looping process comprising of three key components: a power plant, an oxy-fired
CFB power plant as a calciner and a fluidized bed carbonator. The capital cost of the
calciner was estimated based on that for a CFB oxy-fired power plant as both
systems works under the same principle and contain similar units: a CFB oxy-fired
combustor, a steam cycle, an ASU and a CO2 compression unit. Detailed capital cost
estimation for the carbonator was not included and that of this unit was roughly
estimated to be 10% of the total capital requirement of the oxy-fired system.
According to the reported values, the cost of CO2 avoided for the Ca-looping process
was around 15 $/ton CO2 in comparison to 24 $/ton CO2 for a CFB oxy-fired power
plant. It is needed to be highlighted that the given cost estimates reflect the capital
and variable cost data available around the publication year of the reference and
should be adjusted using the cost indices to assess current values. MacKenzie et al.
(2007) estimated a cost for CO2 avoided of 20 $/ton CO2 for the Ca-looping process
while the value of the conventional amine process was around 55 $/ton CO2 (Tuinier
et al., 2011). It was emphasized that the sorbent cost is a crucial parameter in cost
calculations. Thus, when two processes using different sorbents are compared, it is
really important to do this comparison in terms of both energy consumption and cost
of CO2 avoided as it does not mean that the one with less energy consumption will
provide lower cost of CO2 avoided. Romeo et al. (2009) concluded also that the
sorbent costs should be maintained at reasonably low levels to achieve a low cost of
CO2 avoided. A modified sorbent would be preferable only if it provides significantly
better CO2 uptake capacity compared to natural limestone, which may allow a
reduction in purge flow rates.

1.6 Objectives of the Dissertation

Cement is a key construction material. Owing to the increasing demand in rapidly


developing countries such as India, China and Brazil, the demand of cement has
increased by around 6% in 2012 when compared to 2011, reaching 3.78 billion tons
(CW Group, 2012). The cement industry is the second largest stationary CO2

33
emission source in the world and accounts for over 2 billion tonnes of CO2 emitted
per year (IEA, 2009b; CW Group, 2012). Thus, it can be specified as a potential
candidate for the implementation of CCS technology. On the other hand, there is a
growing trend of using sustainably-grown biomass as a renewable energy source in
combustion systems as this is associated with zero net CO2 emissions (Faaij et al.,
1998). The application of CCS to biomass fired systems was reported to produce
negative emissions of CO2 (Ishitani and Johansson, 1996). Therefore, the principal
objective of this dissertation is to evaluate the techno-economic performance of the
Ca-looping process to mitigate CO2 emissions from cement plants and biomass-fired
power plants. This study also contributes to the assessment of various alternative
carbon capture processes including oxy-combustion, amine, indirect calcination and
membrane. In addition, an advanced configuration of the Ca-looping process where
the energy intensive ASU is replaced with a CLC cycle has also been considered as
an option to reduce the energy penalty associated with this system. The primary aims
of research can be given as:

(i) Development of full process simulations: To investigate the process


performance of the carbon capture technologies, the first task was to develop
and simulate full process flowsheets considering all major units having an
influence on mass and energy balances, and chemical reactions in UniSim
Design Suite from Honeywell. The detailed analysis of the base plants was
used for the selection of optimal process configurations and the determination
of a variation in energy consumption when a CO2 capture facility is present.
Technical challenges involved in process arrangements have been addressed,
and innovative solutions have been provided. To be able to predict accurately
CO2 recovery in the carbonator of the Ca-looping process, a rigorous
carbonator model that is not available in the commercial process simulator
has been incorporated into UniSim as a user defined operation.
(ii) Application of an economic analysis: The comparison of different carbon
capture systems should take economic criteria into account. A detailed
economic study based on the assessments of total capital requirement,
operating and maintenance, and variable costs has been carried out to reveal
economics of the proposed carbon capture processes. Since the available

34
methods to estimate the cost of carbon capture has been built on the similarity
of carbon capture systems with commercially available technologies, i.e. the
calciner in the Ca-looping process vs. an oxy-fired power plant, and limited
variable cost data, a sensitivity analysis had to be included to illustrate
variation in current estimates for different scenarios.

1.7 Organization of the Dissertation

This dissertation has been organized into six chapters. A brief overview of the
content of each chapter is as follows:

Chapter 2 describes the technical details of an exemplary dry-feed cement plant


and presents a conceptual design of integrating the Ca-looping process in the cement
plant. It also considers the purge rate of part of the circulating CaO, given the
tendency of the material to sinter and reduce its capture capacity. The purge stream
from the calciner substitutes limestone in raw meal for clinker production. The effect
of molar flowrate ratio of lime make-up to feed CO2 (F0/FCO2) between two
operational limits has been investigated.

In Chapter 3, alternative carbon capture systems have been evaluated for the
purpose of CO2 capture from cement plants. A variety of process configurations to
incorporate a CLC cycle into the Ca-looping process have been initially investigated.
This system allows a reduction in energy penalty regarding to use of oxygen from an
energy intensive ASU. The process integration of an indirect calcination process
analysed in this chapter aims to minimize the thermal energy consumption by using
excess heat from hot gas streams for raw meal preheating. Since the achievable CO2
avoidance rate by the standalone indirect calcination process is only at moderate
levels, an amine process has been added to increase the avoidance rate further.
Besides this, a standalone application of an amine process where a combined heat
and power (CHP) plant provides heat for solvent regeneration has been analysed.
Lastly, two feed gas locations and different multi-stage membrane configurations
have been assessed for the process integration of a membrane separation process.

35
In Chapter 4, the economic performances of the carbon capture technologies
reported in Chapters 2 and 3 have been evaluated. A detailed cost methodology and
resulting levelised cost of cement (LCOC) and cost of CO2 avoided estimations have
been presented for each case. In addition, sensitivities on cost parameters have been
examined. It is the final chapter regarding to CO2 capture from cement industry.

In Chapter 5, the main purpose is to present in-situ CO2 capture by using the Ca-
looping process from a large-scale (>100 MWe) dedicated biomass-fired power plant.
The techno-economic performance of the proposed system has been compared
against that of the alternative biomass-air-fired and biomass-oxy-fired power plants.
A heat exchanger network design has been proposed by conducting a pinch analysis
to recover the maximum amount of excess heat from the high temperature gas and
solid streams available in the Ca-looping process.

Finally, Chapter 6 provides a number of conclusions and recommendations for


future work.

36
Chapter 2 Process Integration of a Ca-looping Carbon
Capture Process in a Cement Plant
2.1 CO2 Emissions from Cement Production

The cement industry accounts for more than 5% of global CO2 emission from
stationary sources amounting to 1.88 Gt CO2/year in 2006 (IEA, 2009). It is foreseen
that the emissions for the cement sector will continue to grow in a parallel with
increasing demand for cement (CW Group, 2012) and reach 4.3 Gt CO2/year by 2050
(WWF, 2008). This is regarded as one of the major industrial carbon emission
sources for which it is worth implementing carbon capture and storage solutions. The
CO2 emissions from cement plants originate from different sources; over 50% of the
emissions result from the calcination of limestone in the raw material while the rest
is generated by fuel combustion (40%) and indirect emissions due to use of
electricity (IEA, 2011). The fuel consumption is significant due to the highly
endothermic calcination reaction and high temperature operation in a kiln (around
1450°C). It has been reported that, in the modern technology, the average energy
consumption in a cement plant has been reduced to around 2.9 GJth/ton clinker
(WBCSD, 2009) and CO2 emission by calcination can be increased to almost 70% by
a more efficient use of the fuel and correspondingly lowering CO2 emissions from
fuel combustion (Rodriguez et al., 2009).

CO2 emissions in the cement production process can be partially reduced by


modifications: improving the process for more efficient use of fuel, replacing fossil
fuels with alternative renewables including waste residues, and mixing clinkers with
mineral additives (Hasanbeigi et al., 2012). Even though these measures can reduce
CO2 emissions resulting from fuel combustion significantly, they cannot tackle the
CO2 emission originating from the calcination reaction. Therefore, it is essential to
deploy a carbon capture technology on cement processes in order to reduce CO2
emissions by more than 90%.

37
2.2 CO2 Capture Technologies for the Cement Industry

Several carbon capture technologies including amine scrubbing (IEA, 2008;


Hassan, 2005), ammonia scrubbing (Dong et al., 2012), oxy-combustion (IEA, 2008;
ECRA, 2009), anti-sublimation (Pan et al., 2013), Ca-looping (Bosoaga et al., 2009;
Vera, 2009; Rodriguez et al., 2012; Stallman, 2013) and indirect calcination
(Rodriguez et al., 2011b) have received great interest to capture CO2 from cement
industry. Cement plants are characterized by a single source of emission (the cooled
flue gas from the preheater), with CO2 concentrations of 15-30 mol%, typically
released to the environment from one or two emission points, after providing heat for
raw material drying. Since a certain percent of CO2 emissions from cement plants
originate from the calcination of carbonated materials, fuel decarbonisation by pre-
combustion capture processes is not effective in reducing the emissions from cement
plants. Thus on-going research has been focused on post-combustion and oxy-
combustion processes as the most conventional capture technologies since they are
considered mature technology or at least ready to be implemented shortly.

In the case of oxy-combustion, oxygen is fed to the burners instead of air. There
are two locations in a cement plant where combustion takes places; the pre-calciner,
where fossil fuel or wastes are burnt to reach the calcination temperature of ~900°C,
and the kiln where cement clinker is produced at 1450°C. Oxy-combustion in the
pre-calciner is suggested as a viable option to capture CO2 from fuel combustion and
raw material calcination, avoiding technical uncertainty of operating the cement kiln
under oxy-combustion conditions (IEA, 2008). Dual preheaters with oxy-combustion
pre-calciner have been designed while raw mill, kiln and clinker cooler operate
conventionally. Part of the CO2-rich flue gases from the pre-calciner is recycled back
and mixed with oxygen to prevent excessive flame temperatures. The CO2 avoidance
rate by oxy-combustion in the pre-calciner is 61% since only part of the CO2 can be
captured. The European Cement Research Academy (ECRA, 2009; ECRA 2012) has
been investigating the operation of the cement kiln as well as the precalciner under
conditions of oxy-combustion for new installed cement plants. The theoretical and
experimental outputs of this new design would allow utilization of oxy-combustion
in the kiln, resulting in higher CO2 avoidance rates. The applicability of advanced

38
oxygen production technologies, such as OTM systems, should be investigated to
reduce energy and economic penalties involved in conventional cryogenic air
separation which would reduce the energy consumption of oxy-combustion further.
However, since the OTM are best coupled with a combustion turbine providing hot
and high pressure air (IEA, 2007), it is reasonable that a combined cycle power plant
would be needed for the OTM integration. Therefore, allocation of additional CO2
emissions from fuel combustion to the cement plant should be accepted, or a
decarbonized fuel should be used in the turbine, bringing about additional cost and
plant complexity.

Post-combustion CO2 capture represents the alternative option for end-of-pipe


CO2 abatement and provides low technical risk. The high CO2 concentration in the
flue gas would make amine-based post-combustion absorption attractive for the
cement industry. However, a significant amount of steam is required for solvent
regeneration and this is expected to have high energy penalties since a separate steam
boiler is needed to supply steam to the solvent regeneration stripper (in case of
capture rates > 80%, only 10 to 50% of the heat required for solvent regeneration can
be recovered from the plant waste heat) (Kuramochi et al., 2012). The steam
generator could generate some electricity with one back pressure turbine as well as
steam (IEA, 2008); however, such a steam cycle design results in very low plant
efficiency. If the separate steam cycle could be built in a way of having such a high
complexity as one found in coal-fired power plants, similar energy penalty can be
expected but no one would build a steam cycle with such a high complexity for the
purpose of carbon capture. The amine-scrubbing (MEA) based post-combustion
method has been proposed to reduce CO2 emissions from a cement plant (Hassan,
2005; IEA, 2008). A coal-fired CHP plant has been designed to provide the steam for
solvent regeneration (IEA, 2008). The flue gases resulting from fuel combustion in
the CHP plant is also fed to the amine process. The need of the FGD and the SCR
units are major limits of this configuration, in which up to 85% of the released CO2
can be captured, corresponding to 74% of CO2 emission avoided.

39
2.2.1 Application of the Ca-looping Process in Cement Plants

It has been argued that the Ca-looping process would have lower net energy
consumption than the amine process since the heat of reaction can be recovered by
generating steam and running a steam cycle due to their relatively higher operating
temperature. The Ca-looping process may be even more advantageous when
integrated with cement plants than those combined with any other industrial plants.
The by-products of the process, i.e. the purge flow from the calciner, which is needed
to maintain sorbent activity, can be used as kiln feed while it must be dumped as
waste or transferred to other sites where a cement plant is present for reuse when
combined with power plants. Since ECRA (2007) listed Ca-looping technologies as
one of the promising capture technologies for cement plants, there have been
conceptual studies which proposed various integration models, such as utilization of
purge flow for the cement clinker and synergy between cement and power plants
(Bosoaga et al., 2009; Naranjo et al., 2011).

Rodriguez et al. (2011b) proposed a way of producing CO2 from the calciner by
indirect heating using hot CaO circulating between calciner and external combustor
instead of oxy-combustion, which is named the indirect calcination process. The
proposed design tackles CO2 emission resulting from limestone calcination only,
which accounts for around 50 - 70% of CO2 emission, and is not effective for CO2
emission relating to fuel combustion. To enhance the performance, it was also
suggested that the hot streams leaving the capture system could be utilized as a heat
source for electricity generation. It should be noted that there is no carbonator in this
design as distinct from the regular Ca-looping configuration so the CO2 that both
external combustor and kiln generate cannot be recovered in the process, which
means that this process would be worth considering only if a moderate level of CO2
capture is adequate. The key operating parameters and performances in this process
are summarized below in Table 2-1.

40
Table 2-1 Summary of process configuration models on integration of Ca-looping process with cement plants.

Authors Rodriguez et al. (2011b) Rodriguez et al. (2012) Romeo et al. (2011)
Hot CaO circulates between calciner and Kiln gas is sent to carbonator for carbon capture. All The industrial symbiosis of cement and power plants through Ca-looping
Type of integration CFB combustor. There is no carbonator in CaCO3 from carbonator and fresh limestone are process has been proposed. The flue gases from both power and cement
the scheme. calcined in one common calciner. plants are fed to the Ca-looping process.
Capacity of 3000 ton cement/day
3000 ton cement/day 3000 ton cement/day
reference plant 500 MWe power plant
Petroleum coke used in CFB combustor Petroleum coke used in cement kiln and calciner
In the power plant, the coal composition [wt%]:
[wt%]: [wt%]:
C: 61.6
C: 82.2 C: 82.2
H: 4.9
H: 3.1 H: 3.1
Fuel (existing plant O: 15.5
O: 0.5 O: 0.5
and calciner) N: 1.2
S: 5.5 S: 5.5
Ash: 6.7
N: 1.9 N: 1.9
Composition Moisture: 10.1
Ash: 0.3 Ash: 0.3
Moisture: 6.5 Moisture: 6.5
No information was given for fuel compositions used in the kiln and
calciner.
18% excess oxygen is used.
Carbonator No carbonator in this system T

Operating temperature - 650°C 650°C


Pressure - 1.0 atm 1.0 atm
F0/FCO2 - 4.5 A purge of 3.2% of the total solid inventory is assumed.

FR/FCO2 - - 4 (fixed)
Ws - - -

u0 - - -
Reactor height - - -
Ca conversion - CaO conversion: 30% -
Capture efficiency in - - 90%
the carbonator
Net CO2 avoidance 33%, it becomes 38% if CO2 emission by 99% 95.3% (for cement plant)
rate extra electricity is excluded.

Gas pressure drop - - -


Model used - There is no a reactor model implemented. There is no a reactor model implemented.
41
Calciner

Operating temperature 937°C 950°C 950°C


Calcination efficiency 100% calcination 100% calcination 100% calcination
O2 content in oxidant Air used 25 vol% O2 -
stream (75 vol% CO2 and H2O)

180 bar/50 bar


A sub-critical steam cycle was chosen. Assumed that the lower temperature limit for energy
600°C/600°C
Steam cycle (120 bar/520°C/520°C) recovery is 150°C. The net thermal efficiency of 33%
A pinch analysis has been conducted to recover maximum amount of
No steam bleeds are performed. is estimated for steam cycle.
surplus energy.
Air separation unit

Electricity - 160 kWh/ton O2 220 kWh/ton O2


consumption
O2 purity - - -
CO2 compression and
purification

Final CO2 pressure 100 bar - -


Final CO2 purity > 95 vol% CO2 - -
Compression train 5 turbo compressors + pump - -
(Compressor efficiency) (75% isentropic efficiency) - -
Overall specific - 100 kWh/ton CO2 -
consumption
Results
Total power demand - 37.0 MWe
The total thermal energy consumption of reference case (power + cement
Power production by 31.6 MWe 41.0 MWe plants) increases 6.7% for the integrated system (power + cement + Ca-
steam cycle looping plants)

Energy demand of the 3.0 GJth/ton cement 2.9 GJth/ton cement The thermal energy requirement in the cement plant decreases by 39.5%
base cement plant since purge CaO can be fed to kiln without calcination.

Energy demand of the 6.1 GJth/ton cement 5.5 GJth/ton cement


base cement plant +
Ca-looping process
42
Given the postulate that the calciner purge can be used as kiln feed, a symbiosis
model of a power plant, a Ca-looping process and a cement plant has been proposed
with its mass and energy balances (Romeo et al., 2011) (see Table 2.1). The flue
gases from both plants are sent to the Ca-looping process, and the CaO purge of this
capture unit is returned to the cement plant, mixed with CaO from fresh raw material,
and used as kiln feed. In this way, the CO2 emission and energy consumption of the
cement plant can be drastically decreased due to reduced load for calcination. The
surplus energy from the capture unit has been utilized to generate electricity by
running a separate steam cycle. A pinch analysis has been conducted in order to
recover maximum energy from the carbonator, solid purge, clean flue gases and CO2
stream before compression. Part of this electricity has been used for CO2
compression and air separation units. The total thermal energy consumption
increases by about 6.7% for the integrated system due to addition of capture unit.
CO2 emission avoided was estimated to be 94% on a basis of total CO2 emission at
both industries by this integration system. It must be highlighted that co-location of
the power plant and the cement plant is needed in these cases, since transporting
large amounts of solids between the two plants over long distances would bring
about additional costs and logistic issues. Furthermore, the heat integration between
the two systems could not be implemented.

A similar assessment has been performed by Romano et al. (2013), who


considered the effect of the actual composition of the purge on the maximum
“substitution rate” of the cement plant raw meal. The large amount of CaSO4 and
ashes from coal combustion in the calciner can limit the maximum amount of Ca-
looping purge that can be used in the cement plant. Such a maximum “substitution
rate” strongly depends on the fuel used in the calciner and on the parameters of the
Ca-looping process.

As a direct integration of Ca-looping with a cement plant, Rodriguez et al. (2012)


investigated two alternative processes (see Table 2.1). One is a retrofit replacing the
existing pre-calciner with an oxy-calciner which can achieve 89% CO2 avoidance
and the other is capturing CO2 from the kiln gas using a carbonator in addition to
capturing CO2 by oxy-calciner to improve avoidance rate up to 99%. As similar with

43
other Ca-looping processes, the surplus energies from high temperature streams are
recovered by an integrated steam cycle. It was reported that the total energy
consumption of cement plant increases from 2.93 GJth/ton cement to 5.45 GJth/ton
cement by retrofit for carbon capture. Nevertheless, the carbonation efficiency of the
solid stream containing CaO with clay minerals was roughly estimated in the
reference study. Furthermore, the CO2-rich gas stream from the calciner was
transferred to the preheater for raw meal heating, which may cause a significant
reduction in CO2 purity if any air-leakage into this unit cannot be prevented. The
design of Ca-looping process in a cement plant can be further improved if the
concerns raised are resolved.

In this chapter, the aim is to present a detailed analysis of a typical cement


manufacturing process and study the reasonable selection of location of the capture
process with respect to process conditions when a Ca-looping process is integrated in
a cement plant. Process simulation includes the implementation of a detailed
carbonator model and its incorporation into a full cement process simulation. Effects
of key operating parameters on Ca-looping process have also been investigated.

2.3 Process Simulation of a Cement Plant (Base Case)

Figure 2-1 shows the block flow diagram of a dry cement process, hereinafter
named the base case configuration. The base case includes all the major units in the
cement plant: raw mill; preheaters; pre-calciner; kiln and cooler. The base case
simulation takes into account key reactions taking place in the process of cement
production. Several auxiliary units, such as crushing and milling of the raw materials,
cement mixing and milling with fly ash and gypsum are not included since their
contribution to the energy balance is not as important as the major units included. In
addition, their operations are not affected by retrofitting the carbon capture units into
the cement plant. It should be highlighted that the base configuration has a separate
pre-calciner upstream of a kiln instead of having a single reactor for calcination and
clinkerization since it is well-known that it can provide a lower energy consumption
and shorter kiln length (IEA, 2008).

44
To
Atmosphere

To
Collected
Atmosphere B/F
Gas Flow Dust
Solid Flow
Coal To
B/F Pet Coke F/D B/F
Primary Atmosphere
Air
Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M Pre-C Kiln Cooler
Meal PHE PHE PHE PHE Clinker

Air Air Air Air


Cooling
Tertiary Air
Air

Figure 2-1 Schematic diagram of a cement plant without a CO2 capture unit (Base Case) (IEA, 2008). Abbreviations: R/M, Raw Mill; B/F, Bag Filter; F/D, Fuel
Drying; PHE, Preheater; Pre-C, Pre-calciner.
45
Table 2-2 Chemical reactions and their standard enthalpies considered in this cement plant simulation
(Taylor, 1990).

Reaction ∆H Reference
(kJ/kg)
CaCO3 (calcite) → CaO + CO2(g) +1,782 CaCO3
AS4H (pyrophyllite) → α-Al2O3 + 4SiO2 (quartz) + H2O(g) +224 AS4H
AS2H2 (kaolinite) → α-Al2O3 + 2SiO2 (quartz) + 2H2O(g) +538 AS2H2
2FeO⋅OH (goethite) → α-Fe2O3 + H2O(g) +254 FeO⋅OH
2CaO + SiO2 (quartz) → ß-C2S -734 C2S
3CaO + SiO2 (quartz) → C3S -495 C3S
3CaO + α-Al2O3 → C3A -27 C3A
4CaO + α-Al2O3 + α-Fe2O3 → C4AF -105 C4AF
S + O2 → SO2 -17,813 S
CaO + SO2 + 0.5O2 → CaO⋅SO3 -7,656 SO2

Table 2-3 Composition of the raw meal fed to the raw mill (Taylor, 1990).
wt %
Calcite 72.5
Quartz 6.0
Pyrophyllite 9.0
Kaolinite 2.4
Goethite 1.8
Moisture 8.0
Sulphur 0.3
Total 100.0

It is crucial to identify the chemical reactions occurring in each unit and determine
their conversion rate in order to have accurate mass and energy balances. Table 2-2
shows the reactions being considered which can be classified into the decomposition
of the raw materials and the clinkerization stages (Taylor, 1990). Given the raw meal
composition in Table 2-3, the approximate chemical composition of the four main
clinker phases (wt%), C3S, C2S, C3A, and C4AF, can be estimated by the Bogue
equation (Bogue, 1929).

𝐶3 𝑆 = 4.0710𝐶𝑎𝑂 − 7.6024𝑆𝑖𝑂2 − 6.7187𝐴𝑙2 𝑂3 − 1.4297𝐹𝑒2 𝑂3 (Eq. 2.1)

𝐶2 𝑆 = −3.0710𝐶𝑎𝑂 + 8.6024𝑆𝑖𝑂2 + 5.06383𝐴𝑙2 𝑂3 + 1.0785𝐹𝑒2 𝑂3 (Eq. 2.2)

𝐶3 𝐴 = 2.6504𝐴𝑙2 𝑂3 − 1.6920𝐹𝑒2 𝑂3 (Eq. 2.3)

𝐶4 𝐴𝐹 = 3.0432𝐹𝑒2 𝑂3 (Eq. 2.4)

46
Table 2-4 Comparison of Bogue equation approximation and the simulation results.
Bogue calculation Simulation
Mineral [wt %] [wt%]
Alite (C3S) 60.6 60.4
Belite (C2S) 17.5 17.2
Tricalcium Aluminate (C3A) 11.0 10.9
Tetracalcium Aluminate (C4AF) 8.6 8.2
Free CaO 0.0 0.0
CaO⋅SO3 2.3 2.5
Ash 0.0 0.8
Total 100.0 100.0

The simulated clinker compositions are in good agreement with those estimated
by the Bogue equation as shown in Table 2-4. It is assumed that 30% of sulphur in
the raw material reacts with oxygen to become SO2 in the first preheater and then
leaves the plant via the raw mill. The remaining sulphur is converted into SO2 in the
pre-calciner and subsequently all the SO2 formed reacts with CaO and oxygen to
form CaO⋅SO3 in the pre-calciner and is included in the clinker product (IEA, 2008).
The raw meal having 8% moisture is dried passing the raw mill where it is heated by
contacting the flue gas leaving the preheaters directly. The flue gas flowrate to be fed
to the raw mill is determined such that both gas and solid streams leave the raw mill
at around 110°C. The gas stream, leaving the pre-calciner at 915°C, is cleaned of the
entrained fine particles passing the four cyclones in series comprising the preheater
where it heats up the raw meal up to 760°C. The solid removal efficiency is assumed
to be 94%, 90%, 85% and 80%, respectively from the preheater stage 1 to 4 (Alsop et
al., 2007). It should be noted that calcination and clay decomposition start to take
place at the 4th preheater with the 10% for calcination and 30% for clay
decomposition referring the phase diagram shown in Figure 2-2 (Taylor, 1990). It is
assumed that entire formed CaO reacts with SiO2 in the 4th preheater stage and is
converted belite (C2S).

The preheated raw meal enters the pre-calciner where 90% of the remaining
calcites are calcined and all clays are decomposed into their constituents, such as
alumina, silica, and ferrite, at the operating temperature of 915°C. The conversion
efficiency is set to 70% for the belite formation in this reactor. As the calcination and
clay decomposition reactions are all endothermic, the pre-calciner is supplied with
the heat generated from coal combustion with the tertiary air heated up to 908°C by

47
the clinker cooler. The flue gas leaving the kiln at 1025°C flows into the pre-calciner
in order to lower the CO2 partial pressure and supply an additional heat source for the
endothermic reaction.

The rotary kiln, where cement clinker is produced by counter-current contact of


the gas and solid streams, has been simulated in three separate units so that the
temperature change along the length can be simulated. The first unit, corresponding
to the solid feed end of the kiln, is simulated as a heat-exchanger to heat the solid
stream from 915°C to 1250°C by its direct contact with the kiln gas flowing in the
opposite direction. Subsequently, the temperature of the solid stream increases up to
1450°C by fuel combustion with primary and secondary air in the reactor (second
unit) in order to calcine the remaining calcite and make all clinkerization reactions
completed. Finally, the kiln product formed in the second unit is cooled to 1370°C
with the incoming secondary air at the solid product end (third unit) (see Appendix
A).

Figure 2-2 Clinker phase diagram (Taylor, 1990).

48
The kiln product is sent to the clinker cooler in which it is cooled to 60°C by
ambient air. Even though there is a potential to burn alternative fuels such as tyres
and biomass fuels in the pre-calciner, coal is selected as a heat provider in a parallel
with the reference (IEA, 2008). The high temperature requirement in the kiln restricts
the fuel flexibility; therefore, pet coke with low ash content is generally combusted
in this reactor. Both fuel streams are dried completely by the hot flue gases from
preheater. The excess air ratio is set as 10% to guarantee complete combustion in
both reactors. The compositions of the fuels used in this study has been constructed
from IEA (2008) report and presented in Table 2-5.

Tables 2-6 and 2-7 show the mass and energy balance around the cement plant
obtained by the base case simulation. Given the raw meal composition, 1.66 kg/s of
raw meal is required to produce 1 kg/s of clinker. The CO2 generation intensity is
around 0.8 ton CO2/ton clinker that is within the range of 0.65 – 0.92 ton CO2/ton
cement given as the average CO2 intensity for cement manufacture (IEA, 2007).
Based on the energy balance, the required thermal energy for unit clinker production
is estimated to be 3.13 MJth/kg clinker. The ratio of heat supply into the pre-calciner
and kiln is maintained at 6 to 4.

Table 2-5 Composition of the fuels fed to the pre-calciner, kiln and calciner (* Data from IEA, 2008).
Precalciner Kiln Calciner
(Coal*) (Pet coke*) (Pet coke)
wt % wt % wt %
Carbon 64.4 85.6 87.2
Hydrogen 4.5 3.5 3.6
Nitrogen 1.4 1.8 1.8
Sulfur 0.9 5.3 3.6
Ash 12.1 0.2 0.2
Oxygen 7.2 1.8 1.8
Moisture 9.5 1.8 1.8
Total 100.0 100.0 100.0

49
Table 2-6 Mass balance of the base case simulation [kg/s].
Mass in Mass out
Raw meal 52.41 Clinker 31.61
Air 99.55 Flue gas
Fuel From fuel drying 4.30
Wet coal to pre-calciner 2.26 From raw mill 75.49
Wet pet-coke to kiln 1.18 Excess Air 44.00
Total in 155.40 Total out 155.40

The overall heat of chemical reactions involved in conversion of raw meal to


clinker is estimated about 178.4 GJth/h (1.57 MJth/kg clinker) by the difference
between enthalpy in and out. The overall heat of reaction is lower in this simulation
than in the reference (Taylor, 1990) (1.76 MJth/kg clinker) since it also takes into
account the heat of the two highly exothermic reactions of sulphur conversion to SO2
and its reaction with CaO.

2.4 Process Simulation of a Ca-looping Process

Different carbonator modelling studies have been published in recent years. These
models have been developed with the support of the experimental results from either
lab or pilot trials of the Ca-looping process as well as those for the CFB
hydrodynamics. Shimizu et al. (1999) proposed a bubbling fluidized bed (BFB)
model for the carbonator, which was based on the Kunii and Levenspiel (K-L) model
(Kunii and Levenspiel, 1991) considering bubble and emulsion regions. This model
was then modified by Abanades et al. (2004) to include the kinetic model proposed
by Bhatia and Perlmutter (1983). Abanades (2002) proposed a simplified carbonator
model that neglects the effects of reactor hydrodynamics. This model only considers
the maximum average carbonation degree of the sorbents based on the maximum
carbonation degree after a number of cycles and the mass fraction of these particles
in the system. The simplified model has been used in many studies for the estimation
of carbonation efficiency (see Appendix B). Later, the model developed by Alonso et
al. (2009) combined the sorbent kinetics and residence time distribution functions.
Although very simple assumptions about the fluid dynamics: plug flow for the gas
phase and the perfect mix of the solids were considered, it has been proven to be
accurate for carbonator design and optimization.

50
Table 2-7 Energy balance of the base case simulation [GJth/hr].

Enthalpy in Enthalpy out


Sensible *Heat by Sensible Heat of
Heat combustion Heat Reaction
Raw Meal 1.82 Clinker 5.12
Air 3.25 Flue gas
Fuel From fuel drying 4.29
Wet coal to pre-calciner 0.12 216.58 From raw Mill 72.95
Wet pet-coke to kiln 0.05 139.25 Excess Air 45.77
Heat lost by radiation and convection 54.54
178.4
Overall heat of reaction
(1.57 MJth/kg)
Total in 361.07 Total out 361.07
The reference state for enthalpy is at 0°C and 101 kPa.
* The heat by combustion of the fuel is a standard heat of combustion at 25°C and 101 kPa.
51
Lasheras et al. (2011) implemented a 1D carbonator model into a full-scale power
plant. The carbonator was modelled as a CFB, and the model was divided into three
main parts: particle distribution in the riser, absorption kinetics and calculation of
overall carbonation efficiency. The particle distribution part was developed for fast
fluidization as given in Kunii and Levenspiel (1991). Two regions were
distinguished inside the carbonator: a lower dense region and an upper lean region.
The rate of carbonation takes the boundary layer diffusion and the spherical grain
model suggested by Abanades et al. (2004) into account. To estimate the carbonation
efficiency, the model for gas conversion in catalytic reactions was used (Kunii and
Levenspiel, 1991).

The model employed for predictions of carbonation efficiency in this dissertation


was developed by Romano (2012). The vital difference between this model and
others is the application of the effects of coal ash and sulphur species to CO2 capture
efficiency. The fast fluidized bed carbonator model is briefly presented here while
further details, including the manner in which it performs integration with a full
process flowsheet, can be found in the original paper (Romano, 2012) and Appendix
B. The model assumes: (i) uniform temperature, (ii) no gas side mass transfer, (iii)
perfect mixing, (iv) uniform particle size and (v) uniform superficial velocity.
Similar to the model proposed by Lasheras et al. (2011), the carbonator model
presented in Romano (2012) was based on the K-L theory for CFB systems. The
reaction model describing the reaction rate of cycled particles was defined as the
following equation (Grasa et al., 2008a):

𝑑𝑋
= 𝑘𝑟 �𝐶𝐶𝑂2 − 𝐶𝐶𝑂2,𝑒𝑞 � = 𝑘𝑠 𝑆𝑁 (1 − 𝑋)2/3 �𝐶𝐶𝑂2 − 𝐶𝐶𝑂2,𝑒𝑞 � (Eq. 2.5)
𝑑𝑡

where X is the carbonation degree, t is time, kr is the kinetic constant, SN is the


specific surface area after each carbonation-calcination cycle, and ks is the instinct
kinetic constant. CCO2 and CCO2,eq refer to actual and equilibrium CO2 concentrations,
respectively. Despite being neglected in most modelling studies, Grasa et al. (2008b)
indicated experimentally that a strong effect on the structure can be detected when
limestone experiences sulphation after each cycle. To reflect the impact of sulphation
on CO2 capture efficiency in the carbonator, the experimental data from Grasa et al.

52
(2008b) was adapted to the model. The experimental data from the reference is fitted
using the following equation and employed in the carbonator model for the
calculation of the maximum carbonation degree of the sorbent.

1
𝑋𝑚𝑎𝑥,𝑁 = + 𝑋𝑟 (Eq. 2.6)
1⁄(1−𝑋𝑟 )+𝑘𝑁

where Xr and k are the constants and N is the cycle number.

The reactor was divided into a dense region with a core-annulus radial distribution
and a lean upper region. The use of the K-L model allows the estimations of solid
distribution, the heights of the bottom dense and upper lean regions as well the mass
velocity of solids. The gaseous phase mass balance was developed by rearranging the
kinetics given in Eq. 2.5 in the form of:

𝑑(𝐶𝐶𝑂2 −𝐶𝐶𝑂2,𝑒𝑞 ) 𝜉 𝑉𝑠 𝜌𝑠,𝑎


=− 𝑘𝑟 �𝐶𝐶𝑂2 − 𝐶𝐶𝑂2,𝑒𝑞 � (Eq. 2.7)
𝑑𝑡 𝑉𝑔 𝑀𝑠,𝑎

where 𝜉 is the volume ratio between potentially active sorbents and the total solids
(including ash and CaSO4). Vs and Vg refer to volumes of solid and gas phase,
respectively. Ms,a and 𝜌𝑠,𝑎 are the molecular weight and density of potentially active
sorbents, respectively. Eq. 2.7 can be solved for the core and wall regions in the
bottom dense section. The final form of the equation representing CO2 concentration
at the dense region outlet is given in Appendix B. The following material balance
was arranged for the lean region.

𝑑(𝐶𝐶𝑂2 −𝐶𝐶𝑂2,𝑒𝑞 )
𝑢0 = −𝜉𝜀𝑠,𝑙 𝜂𝑙 𝑘𝑟𝑖 �𝐶𝐶𝑂2 − 𝐶𝐶𝑂2,𝑒𝑞 � (Eq. 2.8)
𝑑𝑧

Here, u0 is the superficial velocity of gas, 𝜀𝑠,𝑙 is the volumetric solid fraction in the
lean region and kri is the first order kinetic constant of the carbonation reaction. 𝜂𝑙 is
the contact efficiency in the lean region starting from that in the dense region to take
into account the non-ideality of the reactor.

The volume ratio of active solids and the average kinetic constant should be
known in order to find a numerical solution for the mass balance equations. Thus, the
second part of the model was dedicated to the estimation of solid population in the

53
bed, where the impact of ash and CaSO4 contents in feed sorbent on solid residence
is included. Overall, there are two solution steps for the carbonator model. The first
step is related to solid distribution in the reactor and mass balance for the dense and
lean regions, while the second step is dedicated to the estimation of solid residence
time and the average carbonation degree of the sorbent. The numerical solution is
relevant to the calculation of the same CO2 capture efficiency in these two parts.
Despite the fact that the first results for pilot-scale applications have just been
released, the model has provided satisfactory results against those from the lab-scale
facilities operating at Stuttgart University, Germany and INCAR-CSIC, Spain
(Charitos et al., 2011). In all simulations, the operating condition of the carbonator is
carefully chosen to capture 90% CO2 from the feed gas and the CaCO3 fed to the
calciner is regenerated to CaO at complete conversion. The calciner temperature is
selected to be 930°C, which is higher than the pre-calciner temperature (915°C), to
guarantee complete calcination due to the CO2 partial pressure in the calciner close to
1 atm. The temperature of the carbonator, which should be kept as close to the
calciner temperature as possible in order to save the energy consumption for
reheating the circulating solid, is fixed at 650°C in this study.

It is assumed that all SO2 generated by combustion in the calciner is captured by


CaO. As a result of the deterioration of the CO2 absorption capacity through the
carbonation/calcination cycles, fresh CaCO3 needs to be added into the calciner
while same amount of spent sorbents (CaO) are removed from the calciner on molar
basis. The CaCO3 make-up, fuel, and oxygen streams fed to the calciner are neither
preheated nor dried but directly fed to the calciner due to the lack of flue gas
availability for preheating. All the mathematical models for the carbonator are solved
in Matlab and then the carbonator unit was incorporated into the UniSim process
simulation for cement plant as a user defined operation. The Visual Basic code in a
user defined operation transfers the input values from the UniSim into the Matlab
environment where the design calculations are implemented via a component object
model (COM) interface. The calculated values are then sent back to UniSim and used
for the mass and energy balance calculations in the complete process flowsheet (see
Appendix B).

54
(a)

Raw mill 1st Preheater 2nd Preheater 3rd Preheater 4th Preheater Precalciner Kiln Cooler
2000

Temperature (o C) 1500
Gas Temperature
1000

500
Solid Temperature
0

(b)

Raw mill 1st Preheater 2nd Preheater 3rd Preheater 4th Preheater Precalciner Kiln
40
CO2 concentration

30
(mol%)

20

10

Figure 2-3 Variations of (a) gas and solid temperatures and (b) CO2 concentration along the cement
plant.

2.5 Process Integration of a Ca-looping Unit with a Cement Plant

One of the important issues while integrating a Ca-looping unit with a cement
plant is the selection of a feed gas stream for the Ca-looping process. As the flue gas
from pre-calciner flows through the process in the opposite direction to the solid flow
for heat recovery, its temperature and CO2 mol fraction varies over the process as
shown in Figure 2-3. Therefore, the flue gas stream for the capture process should be
selected taking into account the operating condition of a selected capture unit, ease of
heat integration, and CO2 partial pressure.

For the ease of retrofit, it can be envisaged that the flue gas stream after the raw
mill and fuel drying would be an optimal feed for the capture unit. However, the flue
gas at this location has around 22 vol% CO2 as shown in Figure 2-3(b), which is the
lowest value over the entire process. The volumetric flowrate is at its highest at this
point so a larger equipment size of the capture unit would be required. Furthermore,

55
the flue gas may need to be reheated prior to being fed to the carbonator in order to
initiate the carbonation reaction as the temperatures of the flue gases are around
110°C. In addition, the heat of the CO2-depleted gas from the carbonator needs to be
recovered to improve the energy efficiency but an additional facility for heat
recovery should be deployed for this purpose.

In conclusion, the applications of Ca-looping processes to the end-of-the-pipe gas


streams of the cement plants would require a complexity similar to those of power
plants. However, the flue gas at the exit of the 3rd preheater has a temperature of
around 650°C, as shown in Figure 2-3(a), which is similar to the operating
temperature of the carbonator. It indicates that the flue gas from the 3rd preheater
would not require any pre-heating of the feed gas and would be preferable for the
start-up of the carbonator. Moreover, there is no need to recover the heat from the
CO2-depleted flue gas stream for power generation and instead it is possible to return
it back to the cement plant in order to heat up the raw materials in a similar way to
the operation in the conventional cement plant. The flue gas at the 3rd preheater exit
has a higher CO2 concentration (~35 vol%) compared to the end-of-pipe stream (~22
vol%) as shown in Figure 2-3(b) and, in proportion, such a lower gas flow rate would
require smaller carbonator size leading to lower capital expenditure. Therefore, a
decision was made that the flue gas from the 3rd preheater stage is diverted to a Ca-
looping unit for CO2 capture as shown in Figure 2-4. The CO2-depleted flue gas from
the carbonator is routed to the 2nd preheater stage for preheating the raw material
further. It should be noted that it is still possible to capture CO2 from all sources
including calcination and fuel combustion with this configuration, which was
initially proposed by ECRA (2007).

The CO2 depleted flue gas flowing from the carbonator to the 2nd preheater would
have a lower flowrate than that in the base case as a result of carbon capture and its
heat duty is not large enough to heat the raw material up to a temperature that would
be reached in the base case. Thus, part of the excess air from the clinker cooler as
well as CO2-depleted flue gas should be utilised for heating raw material as shown in
Figure 2-4. In all cases of this study, the flowrate of excess air being sent to the raw
mill was determined to heat up the raw material entering the 1st preheater up to

56
110°C. It should be noted that the clay components are not fed into the capture unit
in this study as distinct from the reference (Rodriguez et al., 2012) since the sorbent
performance has not been proven for this mixture yet and there would be a need of
additional efforts for the circulation of inerts. As the purge stream of the Ca-looping
unit is mixed with pre-calcined raw materials and then the mixture is sent to the kiln
for clinker production, the mass and energy balance in the cement plant is
significantly affected by changing the CaCO3 make-up flow rate of the Ca-looping
process. Firstly, since the CaO for clinkerization can be produced in the calciner as
well as the pre-calciner, the ratio of calcite to clay in the raw meal should be
decreased with an increasing F0/FCO2 in order to maintain the same clinker
composition as that in the base case. Subsequently, the decreasing ratio of calcite to
clay in the raw meal results in reduction in heat demand in the pre-calciner and, to a
less extent, kiln. In order to save the energy consumption for a Ca-looping process
further, it is possible to recover the heat of reaction in the carbonator, the heat from
the CO2-rich stream and excess air. The heat of those hot streams can be recovered
by way of generating steam for a steam cycle. The power generated can be utilised
for the cement plant operation, the CO2 compression unit, the ASU, etc.

Since the solid removal efficiency is not 100% on the 3rd preheater stage, a new
cyclone with higher efficiency has been included to prevent the solid transfer from
cement plant to capture unit for precise prediction of the carbonation efficiency in
this unit. It is estimated that the additional pressure increment of the gaseous feed
flowing to carbonator would be approximately 0.16 bar that would be sum of the
pressure loss in operating the carbonator (pressure drop along the carbonator bed,
0.10 bar + gas injection through the nozzle, 0.03 bar) and the pressure drop relating
to the additional cyclone (0.03 bar) (Alsop et al., 2007). 0.20 bar of a total pressure
loss including a 25% safety margin is estimated. The boost of the gas stream pressure
has been made by increasing the cooling air pressure flowing to the cement kiln. The
ASU power consumption is set as 231 kWh per ton O2 product at 99.5% oxygen
purity (DOE, 2003).

57
Excess
Air

Gas Flow Steam CO2 Compressed


ASU Cycle Comp. CO2
Solid Flow
Oxygen
Heat stream
Make-up To
Qcarbonator
CaCO3 Atmosphere
Pet
Coke
B/F Collected
To Dust
Atmosphere
CaCO3
Pet Coke
Carb Cal Coal F/D
CaO
B/F Purge Primary B/F
Air
Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M Pre-C Kiln Cooler
Meal PHE PHE PHE PHE
Clinker

Air Air Air Air

Tertiary Air Cooling


Air

Figure 2-4 Schematic diagram of the proposed process integration of a cement plant with a Ca-looping unit. Abbreviations: R/M, Raw Mill; B/F, Bag Filter; F/D, Fuel
Drying; PHE, Preheater; Pre-C, Pre-calciner; Carb, Carbonator; Calc, Calciner; ASU, Air Separation Unit; CO2 Comp, CO2 Compression.
58
The CO2 compression unit consists of a four-stage turbo compressor with
intermediate cooling, followed by a pump once the CO2 becomes a dense phase. The
inlet temperatures of each stage are fixed at 45°C and adiabatic efficiency of each
compressor is assumed to be 75%. The power requirement for CO2 compression up
to 150 bar is estimated at 1.08 MJth/kg CO2 using a 0.4 conversion factor of power to
equivalent thermal energy. An example mass and energy balance calculations for the
capture cases can be seen in Appendix A.

2.6 Results and Discussion

Two mathematical models for the carbonator have been compared in this study as
shown in Appendix B. While it is assumed that all the active fraction of CaO reacts
with CO2 in the feed gas in the ‘simple model’, the ‘rigorous model’ includes the
effects of both hydrodynamics and reaction kinetics in the fluidised bed reactor.
Moreover, the effects of sulphation on the maximum carbonation degree were taken
into account in the rigorous model based on the experimental data of Piaseck for
limestone sulphation up to 1% in each carbonation/calcinations cycle (Grasa et al.,
2008b). Given a F0/FCO2, the corresponding FR/FCO2 to achieve 90% CO2 capture has
been evaluated using the simple and rigorous models with results shown in Figure 2-
5. The minimum value of the F0/FCO2 being examined is set as 0.20 since the heat
demand at the raw mill cannot be met even by employing the entire excess air in
addition to the flue gas at a F0/FCO2 less than 0.20. The heat requirement in the raw
mill keeps decreasing with increasing F0/FCO2 since the flowrate of raw meal into the
raw mill decreases with an increase of the ratio. The upper limit of the F0/FCO2 ratio
is determined as 5.10 since there is no calcite in the raw meal at this condition, that is
to say, all the calcites in the feed are fed to the calciner. Therefore, the carbonator
captures CO2 generated only from the fuel combustion at this ratio.

59
6 1

5 0.8
Rigorous model (w/ S)
Rigorous model (w/o S)
Simple model

Sulphation Levels [%]


Sulphation levels (%)
FR/FCO2 4 0.6

3 0.4

2 0.2

1 0
0 1 2 3 4 5
F0/FCO2

Figure 2-5 Corresponding FR/FCO2 in range of F0/FCO2 ratio to reach 90% capture efficiency in the
carbonator of proposed configuration (For the rigorous model, the carbonator temperature and
pressure drop along the column were set as 650°C and 0.1 bar, respectively while superficial velocity,
u0, was estimated to be 6 m/s. The sulfidation level is shown at each F0/FCO2).

As shown in Figure 2-5, it is clear that as the FR/FCO2 ratios estimated by the
rigorous model using sulphur-free fuel are definitely higher than those by the simple
model in range of the F0/FCO2 investigated. The extent of difference between the two
models is affected by the residence time of sorbents in the carbonator which is
determined by the amount of sorbent inventory in the reactor. However, when
utilising the fuel having sulphur in the calciner, the required FR/FCO2 ratios need to be
increased way above those with sulphur-free fuel since the CaO is significantly
deactivated by sulphation. In this study the sulphur content in the fuel used in the
calciner was adjusted so that the maximum sulphation, obtained at the 5.10 F0/FCO2
case can be 1%. It implies that the use of sulphur-free fuel would alleviate the
severity of its operation condition due to the lower amount of solid circulation
required given a F0/FCO2 ratio.

60
Figure 2-6(a) shows the variation of the thermal energy requirement per unit
clinker in terms of fuel combustion (pre-calciner, kiln, and calciner) and power
consumption (ASU and blower) with the F0/FCO2. The thermal requirement for the
pre-calciner decreases with increasing F0/FCO2 as the heat requirement for the pre-
calciner decreases in proportion to the reduction in calcite fed to the raw mill. The
thermal requirement for the kiln has decreasing trends with the F0/FCO2 too since it is
assumed that the calcite is completely calcined in the calciner while its conversion is
only 90% in the pre-calciner. However, the reduction of energy demand in the kiln is
not as significant as that in the pre-calciner because the solid flowrate to the kiln are
almost constant regardless of the F0/FCO2 due to nearly constant clinker production
rate in all cases.

The heat requirement for the calciner shows a minimum over the F0/FCO2 range
investigated. Before the minimum, it is decreasing due to decreasing circulating
amount of solid, that is to say, the FR/FCO2 as shown in Figure 2-5. However, after
the minimum, the effect of the increase in the heat duty at the calciner caused by the
F0/FCO2 increase dominates. The total fuel requirement shows a steady overall
decrease with the increase of the F0/FCO2 ratio. The energy requirement for the ASU
is proportional to the fuel consumption in the calciner and that for the cold air blower
is constant with the F0/FCO2 ratio. The electric power consumptions in the ASU and
blower are converted to their corresponding thermal energy consumption using a
power plant efficiency of 0.4. This allows an overall comparison of different options
in terms of the equivalent total thermal energy required. At least, in terms of total
thermal energy consumption for the fuel, the ASU and the blower it is preferable to
operate a Ca-looping process at as high F0/FCO2 ratio as possible.

61
(a)

10

9 Total energy input (Fuel + ASU+Blower)

Thermal energy consumption per unit clinker


Total fuel input
8 Fuel input to calciner
Fuel input to pre-calciner and kiln
[GJth/ton clinker] 7

0
0 1 2 3 4 5
F0/FCO2

(b)

6
Thermal energy per unit clinker [GJth/ton clinker]

Net Energy Consumption


Total Heat Avail
4 Heat Avail from Carbonator
Heat Avail from CO₂-rich Stream
Heat Avail from Excess Air
3 Energy Spent for CO₂ Compression

0
0 1 2 3 4 5
F0/FCO2

Figure 2-6 (a) Energy consumption per unit clinker with respect to fuel and power and (b) net energy
consumption per unit clinker considering heat recovery for power generation.

62
On the other hand, it is intuitively conceivable that a Ca-looping process can be
made more efficient in terms of energy consumption if it is combined with a steam
cycle for heat recovery. There are three different sources from which the heat can be
recovered by generating steam and subsequently running a steam turbine. It is
possible to generate steam by evaporating the water inside the carbonator in order to
keep the reactor temperature constant at 650°C and also recovering the heat from a
CO2-rich stream and the excess air as shown in Figure 2-4. In case of heat recovery
from the two gaseous streams, it is assumed that the hot gas can supply the steam
cycle with thermal energy which is estimated as an enthalpy to be generated when
cooled down to 150°C.

At 0.2 F0/FCO2, as shown in Figure 2-6(b), the heat that can be recovered in the
carbonator is a maximum over the range due to the greatest heat of reaction
generated in the carbonator and the largest amount of hot solids conveyed from
calciner to carbonator (see Figure 2-5). There is no heat to be recovered from the
excess air since all the excess air should be diverted to the raw mill in order to
compensate the deficiency of heat duty of the flue gas. Considering energy
consumption inclusive of CO2 compression, the net energy consumption per unit
clinker production is in the range of 5.2 to 5.5 GJth/ton clinker. This is equivalent to
around 66% increase in energy consumption of a cement plant in producing same
amount of clinker.

A work on preliminary steam cycle design to evaluate the power generation from
the recovered heat has been conducted. The turbine adiabatic efficiencies have been
fixed at 86%, 86% and 95% for HP, IP and LP turbines, respectively (Ahn et al.,
2013). With the support of proposed configuration (see Appendix A), a lumped
conversion factor of 0.44 has been applied throughout this study for rough estimation
of power generation out of the total heat to be recovered from the cement plant
integrated with the Ca-looping process. The estimated power generation is shown in
Figure 2-7. It is predicted that the power generated in steam cycle can exceed the
power demand in the cement plant integrated with a Ca-looping process up to ~1.5
F0/FCO2. The power use for the cement plant operation is assumed as 120 kWh/ton
clinker (IEA, 2008; Taylor, 1990) regardless of the F0/FCO2.

63
0.7

Gross power generation

Power per unit clinker [MWh/ton clinker]


0.6 Total power consumption
CO₂ compressor power

0.5 Cement plant power


ASU+Blower power

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5
F0/FCO2

Figure 2-7 Variation of power generation by heat recovery and power consumption with F0/FCO2 ratio.

The following equations can be used to calculate CO2 intensity, CO2 avoidance
rate and incremental energy consumption when a carbon capture technology is
integrated with a cement plant.

𝐴𝑚𝑜𝑢𝑛𝑡 𝑜𝑓 𝐶𝑂2 𝑒𝑚𝑖𝑡𝑡𝑒𝑑


𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 = (Eq. 2.9)
𝐶𝑙𝑖𝑛𝑘𝑒𝑟 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑖𝑜𝑛 𝑐𝑎𝑝𝑎𝑐𝑖𝑡𝑦

𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦𝑏𝑎𝑠𝑒 𝑐𝑒𝑚𝑒𝑛𝑡 𝑝𝑙𝑎𝑛𝑡 − 𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦𝑐𝑎𝑝𝑡𝑢𝑟𝑒 𝑐𝑎𝑠𝑒


𝐶𝑂2 𝑎𝑣𝑜𝑖𝑑𝑎𝑛𝑐𝑒 𝑟𝑎𝑡𝑒 = (Eq. 2.10)
𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦𝑏𝑎𝑠𝑒 𝑐𝑒𝑚𝑒𝑛𝑡 𝑝𝑙𝑎𝑛𝑡

𝐼𝑛𝑐𝑟𝑒𝑚𝑒𝑛𝑡𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑐𝑜𝑛𝑠𝑢𝑚𝑝𝑡𝑖𝑜𝑛 =


𝑁𝑒𝑡 𝑒𝑛𝑒𝑟𝑔𝑦 𝑐𝑜𝑛𝑠𝑢𝑚𝑝𝑡𝑖𝑜𝑛𝑐𝑎𝑝𝑡𝑢𝑟𝑒 𝑐𝑎𝑠𝑒 − 𝑁𝑒𝑡 𝑒𝑛𝑒𝑟𝑔𝑦 𝑐𝑜𝑛𝑠𝑢𝑚𝑝𝑡𝑖𝑜𝑛𝑏𝑎𝑠𝑒 𝑐𝑒𝑚𝑒𝑛𝑡 𝑝𝑙𝑎𝑛𝑡
(Eq. 2.11)
𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦𝑏𝑎𝑠𝑒 𝑐𝑒𝑚𝑒𝑛𝑡 𝑝𝑙𝑎𝑛𝑡 − 𝐶𝑂2 𝑖𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦𝑐𝑎𝑝𝑡𝑢𝑟𝑒 𝑐𝑎𝑠𝑒

The CO2 emissions in each case can be taken from process simulations while the
clinker production capacity remains almost constant with a value calculated for the
base cement plant. The net energy consumption estimates for the Ca-looping process
inclusive of heat recovery have already been given in Figure 2-6(b).

64
It should be noted that what is initially fixed in this study is not the CO2 avoidance
rate in overall cement process but a CO2 recovery in the carbonator. As shown in
Figure 2-8, therefore, the percentage of CO2 avoidance is as low as 92% at 0.2
F0/FCO2 since most CO2 is captured in the carbonator with 90% capture rate but it
increases up to 99% at 5.10 F0/FCO2 because most CO2 is captured with 100% CO2
recovery in the calciner. It is important to highlight at this point that CO2 avoidance
rate differs from overall CO2 capture rate for the Ca-looping process integrated with
a cement plant because it does not take the capture of additional CO2 releasing by
oxy-combustion in the calciner into account. Although the comparison among
various carbon capture technologies has been based on a fixed CO2 avoided rate in
this dissertation, it is worth to mention that the incremental energy consumption
estimates given in Figure 2-8 can be reduced when it is based on total amount of CO2
captured. Furthermore, additional power generation from surplus heat in the Ca-
looping process does not have any influence on CO2 intensity calculations in Eq. 2.9
as it does when the Ca-looping process is integrated with a power plant.

Another limiting case named ‘oxy-calciner only’ has been simulated such that all
calcites are calcined in the calciner separate from the pre-calciner in a similar way to
the 5.10 F0/FCO2 case but there is no carbonator for carbon capture from the kiln gas.
In this case, the percentage of CO2 avoidance is 90% since this process can capture
CO2 relating to calcinations and fuel combustion in the calciner and cannot capture
CO2 generated by fuel combustion in the kiln. The incremental energy consumption
per CO2 avoided without heat recovery also shows decreasing trends with F0/FCO2
similarly to the total energy input per unit clinker in Figure 2-6(a). It implies that it
would be better to generate CaO by oxy-combustion rather than by the pre-calciner
in the existing plant if no heat recovery system is added. The incremental energy
consumption per CO2 avoided at the ‘oxy-calciner only’ and 5.10 F0/FCO2 cases are
5.1 and 5.5 GJth/ton CO2 avoided respectively without heat recovery. It is thought
that the difference between the two cases (0.4 GJth/ton CO2) can be explained by
additional energy consumption resulting from circulating solids between the
carbonator and the calciner. With heat recovery put in place, the resulting energy
consumption further decreases to 2.5 GJth/ton CO2 avoided for the 5.10 F0/FCO2 case
and 2.3 GJth/ton CO2 avoided for the ‘oxy-calciner’ case.

65
12 100

Incremental energy consumption per CO2 avoided


10 98

Energy Consumption without Heat Recovery


At oxy-calciner

[GJth/ton CO2 avoided]

CO2 avoidance rate [%]


8 Energy Consumption with Heat Recovery 96
At oxy-calciner
Carbon Dioxide Avoidance Rate
At oxy-calciner
6 94

4 92

2 90

0 88
0 1 2 3 4 5
F0/FCO2

Figure 2-8 Variation of CO2 recovery based on CO2 emission at base case and incremental energy
consumption per CO2 avoided with F0/FCO2 ratio.

Table 2-8 shows the quantitative difference of fuel and power consumptions
between the cement plant without capture and those integrated with Ca-looping units
and heat recovery steam cycle at various F0/FCO2 conditions. Table 2-8 also includes
the change of mass flow rates of limestones entering the plant through raw mill and
calciner with the F0/FCO2. Since the proposed cement plants integrated with Ca-
looping units are designed such that they can recover heat contained in the hot excess
air by a steam cycle that would be lost in the conventional cement plant, the net
thermal energy consumption in the cement plants with Ca-looping units are always
lower than that in the conventional cement plant as shown in Table 2-8.

For further analysis, the simulations have been repeated to fix CO2 avoidance rate
at 90%, rather than keeping the CO2 recovery in the carbonator at 90% as presented
in Figure 2-8. The F0/FCO2 ratio is limited to 3.0 in this instance because the CO2
avoidance rate is always greater than 90% at 5.10 F0/FCO2. It is because the ‘oxy-
calciner only’ case that provides the target avoidance rate without a carbonator.
Figure 2-9 presents the incremental energy consumption and CO2 recovery in the
carbonator estimates for the current case.

66
Table 2-8 Detailed constituents of the incremental energy consumption per CO2 avoided considering heat recovery [# Unit: GJth/ton CO2 avoided].

Mass Flow Rates (kg/s)


Clay Limestone .
F0/FCO2 ΔH# ASU# CO2 compression# Cold air blower# Total#
Raw Mill Calciner
0.2* 28.85 9.09 -0.24 1.30 1.75 0.15 2.96
0.3 25.57 12.37 -0.23 1.18 1.65 0.15 2.75
0.4 22.82 15.12 -0.22 1.12 1.59 0.15 2.64
0.6 18.70 19.24 -0.20 1.07 1.54 0.15 2.56
0.8 10.28 15.67 22.27 -0.16 1.06 1.51 0.15 2.56
1.2 11.81 26.13 -0.15 1.04 1.47 0.15 2.51
1.65 8.62 29.32 -0.13 1.04 1.45 0.14 2.50
3.0 4.02 33.92 -0.11 1.04 1.42 0.14 2.49
5.1** 0 37.94 -0.09 1.04 1.41 0.14 2.50
Oxy 0 37.94 -0.08 1.05 1.37 - 2.34
* The lower limit was defined as the heat requirement in the raw mill can not be met below this point
** The upper limit was defined as no calcite is fed to the raw mill at this F0/FCO2 and more calcite would be fed to the cement plant through the oxy-calciner than
required above this point.
67
12 100

Incremental energy consumption per CO2 avoided


90
10 Energy Consumption without Heat Recovery
80

CO2 recovery in the carbonator [%]


Energy Consumption with Heat Recovery
CO₂ recovery in the carbonator

[GJth/ton CO2 avoided]


70
8
60

6 50

40
4
30

20
2
10

0 0
0 1 2 3 4 5
F0/FCO2

Figure 2-9 Variation of carbonation efficiency and incremental energy consumption per CO2 avoided
with F0/FCO2 ratio at 90% CO2 avoidance.

At very low F0/FCO2 ratios in the 90% CO2 avoidance case, the CO2 recovery in
the carbonator needs to be high since only small amount of CaO is provided from the
oxy-calciner. In contrast, it reduces to around 48%, when majority of CaO is
transferred from the capture system. The results for incremental energy consumption
with heat recovery are very similar to those presented in Figure 2-8. Although the
energy requirement in the calciner reduces as a result of a reduction in the amount of
circulated sorbent, the CO2 intensity of the plant increases in parallel so there is not
any significant change in the incremental energy consumption estimates.

In both cases, it is clear that it would be extremely inefficient to operate a Ca-


looping unit at low F0/FCO2 ratio without heat recovery but CO2 can be recovered
with almost constant energy consumption regardless of the F0/FCO2 ratio if a proper
heat recovery is deployed. Moreover, the electricity required to operate a cement
plant integrated with a Ca-looping process can be generated in situ by a steam cycle
attached to the capture unit without any external source of electricity which would be
associated with carbon emissions.

68
2.7 Concluding Remarks

A way of capturing CO2 from cement plants by integrating it with a Ca-looping


process has been investigated. The cement process simulation implemented in this
study was proven to be reliable in that the total energy consumption estimated by the
simulation lies within the range of those reported in the literature and the clinker
compositions estimated in the simulation are in good agreement with those calculated
by the Bogue equation. Among the flue gas streams, the gas stream leaving the 3rd
preheater was selected to be the optimal feed suitable for the Ca-looping capture unit
since 1) it does not have to be preheated, 2) it has a higher CO2 partial pressure and a
lower total volumetric flowrate, and 3) a simpler design of the steam cycle for heat
recovery is possible.

The upper and lower limits of the F0/FCO2 ratio have been set in order to see the
effect of F0/FCO2 on the energy consumption. Given 90% carbon capture in the
carbonator, the CO2 avoidance rate ranges from 92% to 99% depending on the
F0/FCO2 ratio. The incremental energy consumption by carbon capture decreases with
the F0/FCO2 ratio, but with heat recovery from the capture unit, the energy
consumption can be almost constant regardless of the ratio. It was observed that the
incremental energy consumption estimates remain almost constant even if the CO2
avoidance rate is fixed at 90%.

It should be noted that there may be a constraint in the minimum fuel supply to
the kiln to ensure a stable operation in the kiln as pointed out in the IEA study
(2008). Therefore, the actual upper limit of the F0/FCO2 ratio needs to be defined
considering plant operability. Moreover, the estimation of the amount of heat that
can be recovered from three high temperature sources can be made more accurate by
inclusion of a detailed steam cycle in the integrated process flowsheet.

69
Chapter 3 Alternative CO2 Capture Technologies for
Cement Plants

This chapter is divided into three main sections. Section 3.1 introduces an
advanced system combining the Ca-looping process with the chemical looping
combustion (CLC) cycle. The process integration of the indirect calcination process
is explored in Section 3.2. Furthermore, a comprehensive analysis of the standalone
amine process and a hybrid configuration is performed. In Section 3.3, two feed gas
locations and several dual-configurations are assessed to investigate the potential of
incorporating a membrane separation process.

3.1 Combining Chemical Looping Combustion and Calcium


.Looping Process

As discussed in Chapter 2, one of the major energy penalties in the Ca-looping


process results from the need of an air separation unit (ASU) for the production of
high purity CO2 in the calciner, which also escalates the total capital cost
requirement. A novel alternative for the transfer of pure oxygen into a calciner was
proposed by Abanades and Murillo (2009). In their work, a CLC system was coupled
with a Ca-looping process to provide heat to the calciner by the exothermic reduction
of a metal oxide (CuO) as an oxygen carrier with methane.

The schematic diagram of the conventional CLC process is given in Figure 3-1.
This system is comprised of an air reactor and a fuel reactor. In the fuel reactor,
oxygen from the carrier oxidizes the fuel while the depleted oxygen carrier is
regenerated with air in the air reactor. The air reactor always produces heat while the
fuel reactor either requires heat or releases heat depending on the types of oxygen
carrier and fuel (Adanez et al., 2012). While the reduction of majority of the metal
oxides used in the CLC system is endothermic, only a few metal oxides have
exothermic reduction reaction, for example, the reactions between CuO and CH4 or
CO or H2, NiO and CO or H2, and Mn2O3 and CH4 or CO or H2 are exothermic.

70
(N2, O2) (CO2, H2O)

(MexOy-1)
Air Reactor Fuel Reactor
(MexOy)

Air Fuel

Figure 3-1 Schematic diagram of the conventional CLC process.

In the case of integrating CLC with the Ca-looping process, it is important to


satisfy the heat requirement in the fuel reactor (or calciner) because of the limestone
calcination. CuO has been defined as the most promising oxygen carrier for this
purpose in different studies owing to its highly exothermic reduction with a fuel and
high oxygen carrying capacity (Abanades and Murillo, 2009; Abanades et al., 2010;
Manovic and Anthony, 2011b; Manovic et al., 2011c). Therefore, there is a heat
release in both reactors when CuO is selected as an oxygen carrier in the CLC
process. By this way, the exothermic reduction of CuO can provide heat for the
endothermic calcination reaction in the calciner. Besides, it is possible to evaporate
the water inside the air reactor to control the operating temperature which can
increase due to the exothermic oxidation reaction. The heat of fuel combustion in this
process is equal to the total heat release in both reactors.

To date, variety of experimental demonstrations using CuO/CaO sorbents have


been reported for pre-combustion and post-combustion CO2 capture applications
(Abanades et al, 2010; Manovic and Anthony, 2011d). Al2O3 as a support is often
included because of the low melting point of CuO (1085°C) and activity loss of CaO
due to sintering. Abanades et al. (2010), Fernandez et al. (2012) and Martinez et al.
(2014) investigated this combined process for hydrogen production and/or electricity
generation from natural gas by sorption enhanced reforming. Abanades and Murillo
(2009), and Manovic and Anthony (2011d) proposed different process schemes for
its practical implementation for post-combustion applications. Kierzkowska and

71
Muller (2012) reported that a Ca-Cu composite without any inert support can be
manufactured by using a co-precipitation technique. Later, this group indicated that
sol-gel derived, calcium-based, copper-functionalised CO2 sorbents possess excellent
oxygen-carrying and stabilised CO2 uptake capacities (Kierzkowska and Muller,
2013). The performance of MgO supported CaO/CuO (Qin et al., 2012) as well as
the effects of thermal pre-treatment of CuO and steam addition on the sorbent
performance was also examined (Qin et al., 2013).

The main objective of this section is to investigate the potential combination of


the Ca-looping and the CLC processes in order to reduce the energy penalty
associated with the use of an ASU in the process scheme presented in Figure 2-4.

3.1.1 Preliminary Analysis

A way of integrating the CLC into the Ca-looping process, hereinafter called Ca-
Cu looping process, including three different solid routes is shown in Figure 3-2.
This system contains three reactors: a carbonator, a calciner (fuel reactor) and an air
reactor. The solids in the system can be circulated in a direction of carbonator →
calciner → air reactor → carbonator (Route 1). In addition, Manovic and Anthony
(2011d) proposed two other solid circulation routes that are carbonator → calciner →
carbonator, labelled as Route 2 and carbonator → air reactor → calciner →
carbonator (Route 3) as a reverse of Route 1. In Route 2, an air reactor is not
included since it assumes that sufficient oxygen is always present in feed gas to
regenerate depleted oxygen carrier. Nevertheless, the oxygen content of cement flue
gases (around 1 mol% from 3rd preheater exit) is not enough to oxidize large amounts
of oxygen carrier. Therefore, Route 2 has been eliminated. Route 3 is another
interesting option and would be a potential candidate. However, the main concern in
this scheme is the temperature of the air reactor that needs to be strictly controlled to
prevent partial calcination in this reactor. Furthermore, the operating temperature
will be lower than that of the calciner which makes the transfer of heat from the air
reactor to the calciner unfeasible. Hence, only Route 1 is further analyzed in the rest
of this study.

72
Route 1 O2-depleted air
Route 2
Route 3

Air Reactor Air

CO2-depleted
gas CO2-rich gas

Carbonator Calciner

Flue gases Fuel

Figure 3-2 Schematic diagram of the Ca-Cu looping process (Manovic and Anthony, 2011d).

It should be noted that the main property of a metal oxide to be employed in this
system is its exothermic reduction reaction. In this study a CuO/CaO sorbent
supported by Al2O3 is used owing to the benefits explained above. Also, methane is
used as the fuel, but syngas (CO + H2) can also be explored as an alternative fuel for
the reduction/calcination step. While CaO is the CO2 capture agent in the carbonator,
the heat requirement for the calcination reaction can be satisfied by the reduction of
CuO with methane. The Cu leaving the calciner is oxidized back to CuO in the air
reactor. All reactions defined in the process simulations are presented in Table 3-1.
The sum of oxidation and reduction reactions occurring in two different reactors
gives the overall methane combustion reaction as presented below.

𝐶𝐻4(𝑔) + 2𝑂2 (𝑔) → 𝐶𝑂2(𝑔) + 2𝐻2 𝑂(𝑔) ∆𝐻 = −798 𝑘𝐽/𝑚𝑜𝑙 (Rn. 3.1)

73
Table 3-1 Chemical reactions defined in the process simulations.

Reaction Reactor ∆H25°C


(kJ/mol)

CaO(s) + CO2(g) → CaCO3(s) Carbonator −179

4CuO(s) + CH4(g) → 4Cu(s) + CO2(g) + 2H2O(g) −158


Calciner
CaCO3(s) → CaO(s) + CO2(g) +179
2Cu(s) + O2(g) → 2CuO(s) Air reactor −320

It should be highlighted that only approximately 20% of the heat of methane


combustion is usable in the calciner while the remaining should be somehow
recovered in the air reactor. Since the main aim is to provide heat into the calciner
with less expense in the air reactor, it is clear that a method of heat transfer between
the calciner and air reactor is necessary in order to prevent excessive thermal energy
requirement in the system. Otherwise, severe heat duties in the capture plant will be
required, and there is a strong possibility that those values would be much higher
compared to the requirement in the base cement plant.

Rodriguez et al. (2011b) proposed the indirect calcination process that uses high
temperature solid circulation from a CFB combustor to a fluidized bed calciner for
the purpose of transferring heat required for limestone calcination. In the process, the
combustor operates at higher temperatures than the calciner, and heat transferred by
hot CaO particles from this reactor satisfies the heat requirement in the calciner. In
this manner, the air reactor can be operated at a higher temperature than that of the
calciner in Route 1. The surplus heat from this reactor can reduce the heat
requirement in the calciner. It would also affect the sorbent composition as well since
the required methane and CuO flows reduce parallel with a decrease in the calciner
heat requirement. Therefore, the fraction of CuO in the sorbent can potentially be
decreased. Another advantage of this approach would be the elimination of cooling
requirement in the air reactor. While higher reactor temperatures facilitate heat
transfer by reducing solid circulation rates, it is also well-known that they also
trigger sorbent degradation due to the sintering.

74
The air reactor temperature is limited to 950°C, and the effects of high
temperature on sorbent performance in the given conditions have been evaluated by
experimental analyses. To determine the flow rate of air into the air reactor, the O2
molar fraction in the O2-depleted stream is fixed at 3 mol% and complete oxidation
is assumed in each cycle (de Diego et al., 2004; Garcia-Labiano et al., 2004). Since
the reduction of methane with CuO produces only CO2 and a significant amount of
H2O lowering the partial pressure of CO2 in the calciner, the operating temperature
of the calciner is set to 880°C by considering a 15°C increase on the equilibrium
temperature for complete calcination. The excess CuO ratio is fixed at 30% to
guarantee complete reduction of methane in the calciner (Forero et al., 2011). The
carbonator operates at 650°C in the previous chapter but the carbonator temperature
is set to 700°C in this chapter since it allows the reduction of heat duty in the calciner
along with the quantity of solid circulation between the reactors. Although it is well-
known that the higher carbonation temperatures reduce the equilibrium carbonation
efficiency, this assumption also alleviate the heat requirement in the raw mill at very
low F0/FCO2 ratios since the temperature of the CO2-depleted gas stream sent back to
the cement plant will be greater. The rigorous carbonator model has been modified
and used for the estimation of the carbonation efficiency.

Due to the expected degradation of the CO2 uptake capacity as a result of high
reactor temperatures, part of the spent sorbent needs to be replaced with fresh
sorbent. To capture the CO2 resulting from calcination of the fresh sorbent, it should
be fed to the calciner. If the purge stream is removed from the calciner, the heat
released from the Cu oxidation cannot be recovered. To prevent such heat losses, the
purge stream should be removed either from the air reactor or the carbonator. At this
point, it is not clear in the literature how a purge stream containing CaO/CaCO3,
CuO and Al2O3 can be utilized. However, CaO and Al2O3 are cement raw materials
and can potentially be used for clinker production. Kolovos et al. (2005) indicated
that the addition of 1 wt% CuO in cement raw meal promotes sintering and improves
the burnability of the cement raw meal. Also, its favouring effect on the cement
strength development and negligible effect on the physical properties were reported.
Ma et al. (2010) mixed a reference cement raw mix with up to 3 wt% CuO. It was
demonstrated that CuO promotes CaO consumption and improves the clinkerization

75
process. CuO content over 1 wt% in the raw meal caused a decrease in the 3-day and
28-day strengths. Thus it is assumed in this study that the purge stream can be
removed after the air reactor and used for clinker production with a maximum limit
of 1 wt% CuO in the raw meal. However, further studies on this subject is still
required since it was reported in another reference (Bhatty et al., 2011) that the
copper (I) oxide (Cu2O) formed under reducing conditions adversely affects the
formation of alite and belite phases and significantly decreases the fusion
temperature. The use of spent sorbent including CuO would be beneficial for the
reduction of the heat requirement in the base cement plant by eliminating the partial
need for the limestone calcination, but the operating cost of the system will increase
significantly because of the excessive cost of CuO.

It would be preferable to separate the CLC sorbent from the purge and reuse it in
the capture system if CuO shows stable oxidation/reduction performances. However,
this option requires additional investigation, especially for the separation of the CLC
sorbent from the purge and is not considered in the scope of this study. In case of a
possibility of the separation of CuO from the purge stream, it would mainly affect the
economic performance rather than the process efficiency presented in this section.

3.1.2 Experimental Data for the CLC Sorbent

The experimental CaO and CuO conversion data for the CLC sorbent were
provided by Dr. Agnieszka Kierzkowska and Prof. Christoph Müller from the
Laboratory of Energy Science and Engineering at ETH Zurich University,
Switzerland. The sorbent is a physical mixture of Al2O3-stabilized CuO and Al2O3-
stabilized CaO where the material compositions are 87 wt% CuO and 81 wt% CaO,
respectively. The CaO-based sorbent was derived by sol-gel technique while the
CuO-based CLC material was co-precipitation based. The details of manufacturing
these sorbents were already given in the relevant references (Broda et al., 2012;
Imtiaz et al., 2012) and are not repeated here. The CaO and CuO conversion
performances of the sorbent were measured by a thermo-gravimetric analysis (TGA).
The sorbent was calcined at 850°C under 10% CH4 for 20 minutes and carbonated at
700°C with 36 mol% CO2 for 40 minutes. The oxidation took place at 950°C under

76
10% O2 for 25 minutes. It should be noted that part of the experimental conditions
differ from the simulation inputs, for example air is selected as an oxidation agent in
the process simulations and the cycle times are very long for the carbonator model in
use. Furthermore, because of the operational limitations of the TGA system, only
10% CH4 was fed to the system instead of 100% CH4 assumed in the simulations.
The use of diluted CH4 reduces the effect of sintering related to the CO2 partial
pressure. However, it is acknowledged that the currently available data is sufficient
for the preliminary analysis of the proposed configuration with a sensitivity analysis
on the sorbent performance.

3.1.3 Modification of the Rigorous Carbonator Model and Process


.Integration

The following equation proposed by Li et al (2008) was used to fit the


experimental CaO conversion data for the sorbent as it provides an accurate
regression of the data:

𝑋𝑚𝑎𝑥,𝑁 = 𝑎1 𝑓1𝑁+1 + 𝑎2 𝑓2𝑁+1 + 𝑏 (Eq. 3.1)

where a1, a2, f1, f2 and b are the constants, and N is the cycle number.

Another important term that needs to be revised in the rigorous carbonator model,
when it is used in the Ca-Cu looping process, is the mass fraction of particles after N
cycles, rN (Eq. B3 in Appendix B). The solids leaving the calciner are fed to the air
reactor while the majority of the solid stream leaving this reactor is recycled to the
calciner for heat transfer in the Ca-Cu looping process. The effect of high
temperatures on sorbent performance in the air reactor has been included in the
model by implementing relevant experimental data; however, it needs to be kept in
mind that the solid circulation does not exist in the experimental setup. According to
the experiments conducted by Grasa and Abanades (2006) where the calcination and
carbonation temperatures were set to 950°C and 650°C, respectively using La Blanca
limestone, the calcination time is only effective on the sorbent performance for initial
cycles, but this effect disappears in the following cycles. Following this argument,
we assumed that the circulation of solids between the air reactor and the calciner

77
corresponds to having longer solid residence times in the calciner since no
carbonation occurs in the air reactor. Therefore, Eq. B3 has been kept as it is in the
model and the potential effects on the initial performance of the sorbent have been
neglected. Nevertheless, it should be noted that the accuracy of this assumption
diminishes at high F0/FCO2 ratios because the early performance is more crucial if the
sorbent does not stay in the system for a high number of cycles.

The final implementation of the Ca-Cu looping process with the base cement plant
is presented in Figure 3-3. Following the proposed process integration of the Ca-
looping process with the base cement plant given in Chapter 2, the flue gases from
the 3rd preheating stage are diverted to the carbonator for CO2 capture and the CO2-
depleted stream from this reactor is routed back to the 2nd preheater for raw meal
preheating. The heat duty of the CO2-depleted stream is not enough to satisfy the
heat requirement in the raw mill. Thus, the excess air from the clinker cooler should
be used for raw material heating as presented. The majority of excess heat from high
temperature O2-depleted gases leaving the air reactor is transferred to the air feed so
there is a need for a tubular regenerative air heater (DOE, 2003). The temperature of
the O2-depleted stream can be further reduced by preheating the make-up and
methane streams with a final temperature of 150°C. The purge stream containing
CaO, CuO and Al2O3 from the air reactor is mixed with the pre-calcined raw meal
and used for clinker production. Since part of the requirements of CaO and Al2O3 are
provided from the capture unit, the flow rates of these materials in the cement raw
meal have been adjusted in order to keep the clinker production rate similar to that in
the base cement plant.

For a direct comparison of the Ca-Cu looping process with the Ca-looping
process, in addition to the outcomes of Chapter 2, the process scheme presented in
Figure 2-4 has been updated using methane as fuel in the calciner at similar F0/FCO2
ratios determined for the Ca-Cu looping process. There are some benefits of using
methane in this scheme compared to a coal-fired calciner even though the cost of
coal is usually cheaper than that of the methane. First, since a significant amount of
H2O is formed during combustion in this reactor lowering the CO2 partial pressure,
the required calcination temperature for complete calcination reduces. Furthermore,

78
the negative effects of sulphur and ash from coal on the sorbent performance can be
prevented. To be consistent with the Ca-Cu looping process, the calciner also
operates at 880°C in this scheme while the carbonator temperature is set to 700°C.
The experimental data for the natural limestone was also provided by the Laboratory
of Energy Science and Engineering at ETH Zurich University where the limestone
was calcined in 63 mol% CO2 at 880°C and carbonated at 700°C with 35 mol% CO2
during the TGA experiments.

79
Steam CO2 Compressed
Cycle Comp CO2
Make-up
Air Methane
Gas Flow
Solid Flow To
Heat stream Atmosphere To
Qcarbonator Atmosphere
A/R Collected
B/F
Purge Dust
To
Atmosphere
Coal
Carb Cal Pet Coke F/D
B/F Primary B/F
Air

Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M Pre-C Kiln Cooler
Meal PHE PHE PHE PHE
Clinker

Air Air Air Air


Cooling
Tertiary Air Air

Figure 3-3 Schematic diagram of the process integration of a cement plant with an integrated Ca-looping/CLC unit. Abbreviations: R/M, Raw Mill; B/F, Bag Filter;
F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; Carb, Carbonator; Cal, Calciner (Fuel Reactor); A/R, Air Reactor; CO2 Comp, CO2 Compression.
80
1 1

0.8 0.8

CaO conversion (Xmax,N)

CuO Conversion
0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 20 40 60 80 100
Number of cycles

CLC Sorbent Natural Limestone

Figure 3-4 Experimental data for the CLC sorbent (74.0 wt% CaO, 7.5 wt% CuO balanced with
Al2O3) and natural limestone used in the rigorous carbonator model (Ozcan et al., 2013).

3.1.4 Results and Discussion

The experimental performances of the CLC sorbent with a composition of 74.0


wt% CaO and 7.5 wt% CuO balanced with Al2O3 and natural limestone are presented
in Figure 3-4 (Ozcan et al., 2013). With a support of preliminary mass and energy
balance calculations with simultaneous experimental investigations, the composition
of the CLC sorbent estimated by the process simulator (75.9 wt% CaO, 5.5 wt% CuO
balanced with Al2O) is close to that of the experimentally tested CLC sorbent. The
CaO conversion efficiency is initially higher for the CLC sorbent, but its degradation
is more severe compared to the natural limestone, which can be linked to the higher
temperature in the air reactor and corresponding sintering effect. The negative effects
of sintering can be reduced by operating the air reactor at temperatures lower than
950°C (and above 880°C); however, very low temperature differences would
complicate the solid circulation between the air reactor and calciner. The CaO
conversion curves were fitted using Eq. 3.1, and the fitting constants are given in
Table 3-2. For the CLC sorbent, a two-step fitting procedure has been followed for
more accurate regression of the data.

81
Table 3-2 The fitting parameters for the CLC sorbent and limestone.

a1 a2 f1 f2 b
Limestone 0.3836 0.3993 0.8714 0.4437 0.2274
CLC (N≤19) −25.94 0.7126 0.04966 0.929 0.1801
CLC (N>19) 1.072 2.968 0.9306 0.3252 0.1006

The first set of parameters belongs to cycle numbers ≤19 and the second is for the
remaining cycles (up to 24 cycles). It would be preferable to fit the experimental data
for the CLC sorbent when its CaO conversion rate becomes constant but the current
method is appropriate given the data available. After a rapid increase in CuO
conversion rates of the CLC sorbent in the first 4 cycles, it remains stable during the
24 cycle TGA analysis, which proves that there is a strong potential of using purged
CuO again in the capture system. This can also be supported by previous
experimental studies on the subject (Qin et al., 2012; Kierzkowska and Muller,
2013). As mentioned previously, the reuse of purged CuO will allow significant
reductions in the variable cost of the system compared to the case where it is dumped
as waste.

The main variables from mass and energy balances including those for the base
cement plant are presented in Table 3-3. Two different systems, Ca-looping and Ca-
Cu looping processes have been compared, and the impact of the F0/FCO2 ratio has
been investigated. The F0/FCO2 ratio is initially assigned and the required FR/FCO2
ratio is calculated to achieve 90% CO2 avoidance by keeping the CO2 capture
efficiency between 88 – 90% in the carbonator and using an assumption of 100%
capture efficiency in the calciner. The F0/FCO2 ratio is limited to 0.15 to produce a
CuO weight fraction of 1% in the cement raw meal. A smaller value of 0.02 F0/FCO2
has been included to investigate the effect of purge flow on the process performance.
It should be noted that the minimum allowable F0/FCO2 ratio given in Chapter 2 was
determined to be 0.20 whereas it can be reduced to 0.15 here since the carbonator
operates at 700°C so the CO2-depleted stream sent to the cement plant is at a higher
temperature. Nevertheless, for the 0.02 F0/FCO2 case, the excess air stream fed to the
raw mill should be heated up using the hot gas streams available in the capture plant
before being sent to the raw mill, which increases the complexity of the system.

82
Table 3-3 Simulation outputs of the proposed schemes in Section 3.1, including those for the base cement plant.

Base Cement Ca-looping Ca-looping Ca-Cu looping Ca-Cu looping


Plant (I) (II) (I) (II)
Carbonator model inputs
F0/FCO2 - 0.02 0.15 0.02 0.15
FR/FCO2 - 4.2 3.1 9.4 2.0
CO2 recovery in the carbonator (%) - 90.0 88.2 89.7 88.0
CO2 intensity (ton CO2/ton clinker) 0.8 0.08 0.08 0.08 0.08
Thermal energy requirements
(GJth/ton clinker)
Cement plant 3.13 3.09 2.72 3.08 2.67
Capture plant - 4.30 4.0 7.37 4.38
Power generation (MWe)
CO2-rich gas - 13.1 13.5 19.0 16.2
Carbonator - 46.4 36.3 80.4 37.6
Power consumption (MWe)
ASU - 10.0 9.2 - -
Compression (inc. blowers) - 14.7 14.0 18.0 14.7
Cement plant auxiliaries 13.7 13.7 13.6 13.7 13.8
Net power (MWe) −13.7 21.1 13.0 67.7 25.3
CO2 avoidance rate (%) - 90.0 90.0 90.0 90.0
Incremental energy consumption
(GJth/ton CO2 avoided) - 2.6 2.5 2.2 1.7
83
8

6
Limestone (simple model)
CLC Sorbent (simple model)

5
FR/FCO2

1
0 1 2 3 4 5
F0/FCO2

Figure 3-5 Corresponding FR/FCO2 values estimated by the simple carbonator model for the given
range of F0/FCO2 to reach 90% capture efficiency in the carbonator. The results are shown for two
different sorbents, CLC sorbent and natural limestone.

In both configurations, the required FR/FCO2 ratios decrease with increasing


F0/FCO2 rates as more fresh sorbent enters into the system and improves the activity
of the CO2 sorbent. The level of reduction is greater for the Ca-Cu looping process,
mostly because of the change in the average carbonation efficiency (Xave) (see
Appendix B). As the simple model is only based on average carbonation efficiency
calculations, the FR/FCO2 estimates for a predefined range of F0/FCO2 by using this
model is shown in Figure 3-5 to support the results obtained by the rigorous model.
The required FR/FCO2 is greater for the CLC sorbent when F0/FCO2 is smaller than 0.1,
but after this point an opposite behaviour is observed. The difference in the
behaviour can be related to change in sorbent performance after number of
carbonation/calcination cycles as presented in Figure 3-4.

The thermal energy requirement in the cement plant reduces in the capture cases
owing to the transfer of pre-calcined limestone from the capture plant, lowering the
heat duty in the pre-calciner and kiln. With increasing purge flows, the thermal
energy requirement of the cement plant reduces to a minimum of 2.7 GJth/ton clinker.
The modest difference between the 0.15 F0/FCO2 cases can be associated with Al2O3

84
in the purge of the Ca-Cu looping process as this, when supplied from the capture
system, reduces the heat duty associated with its preheating in the cement plant. For
all capture cases, the total thermal energy requirement is more than double of that of
the base cement plant. It is clear that if the heat transfer between the air reactor and
calciner does not exist in the Ca-Cu looping process, the current estimates for this
system would further increase significantly.

Part of the energy can be recovered from the carbonator and CO2-rich gas stream
to drive a steam cycle for power generation, which can then be used to meet the
demand in the cement and capture plants arising from the cement plant auxiliaries,
CO2 compression unit and ASU, if needed. For this purpose, the methodology
explained in Chapter 2 has been retained for the estimation of power generation, and
the calculated power generation/consumption values are also summarized in Table 3-
3. The difference between the power generation and total power consumption is
referred to as net power generation. A negative value was calculated for the base
cement plant since the demand in this plant can only be fulfilled by power import.
The major advantage of eliminating the need for an ASU is the reduced power
consumption in the Ca-Cu looping process. The incremental energy consumption
estimates covering also the power requirement for CO2 compression and ASU (if
needed) and considering heat recovery for power generation reduce to 1.7 GJth/ton
CO2 avoided at 0.15 F0/FCO2 compared with 2.5 GJth/ton CO2 for the Ca-looping
scheme at the same F0/FCO2. In comparison with the outcomes of Chapter 2, where
the effect of sulphation on the sorbent performance is considered, the incremental
energy consumption estimates for the Ca-looping scheme presented in this chapter
are smaller at low F0/FCO2 ratios. This can be linked to the reduced total thermal
energy requirements as a result of the improved performance of the sorbent.

85
140

120 Thermal Energy into Capture Plant

Change in Parameters [%]


Net Power
100
Incremental Energy Consumption
80

60

40

20

0
-60 -50 -40 -30 -20 -10 0
Change in Sorbent Performance [%]

Figure 3-6 The sensitivities of the thermal energy requirement in the capture plant (MWth), net power
capacity (MWe) and incremental energy consumption (GJth/ton CO2 avoided) against the change in
sorbent performance for the Ca-Cu looping (II) case.

The sensitivity of the main results for the Ca-Cu looping (II) case against the
sorbent performance is given in Figure 3-6. Even though it is not possible to predict
precisely the CaO conversion behaviour of a sorbent without conducting relevant
experiments, the CaO conversion performance of the CLC sorbent has been reduced
by 25% and 50% for a brief analysis. The ratio of F0/FCO2 has been retained at 0.15
so the main difference results from the change in required FR/FCO2 to achieve 90%
CO2 avoidance. By lowering the sorbent performance by 25% and 50%, the required
FR/FCO2 increases to 3.3 and 9.1, respectively. At very high FR/FCO2 ratios, the
amount of solid circulation between the reactors increases, and this triggers the
thermal energy requirement in the capture plant. With a reduction of 50% in the
sorbent performance, the thermal energy requirement increases by 55% while the net
power capacity correspondingly reaches 55.5 MWe. The incremental energy
consumption estimate is calculated to be 2.0 GJth/ton CO2 avoided for the worst
scenario in proportion to severe thermal energy requirement and compression unit
duty in the capture plant, but it is still lower than that of equivalent Ca-looping
process for which the results have been given in Table 3-3.

86
Given the postulate that the limitation of 0.15 F0/FCO2 is only valid for the Ca-Cu
looping process to restrict the quantity of CuO in the cement raw meal, this ratio can
be set higher for the Ca-looping process as already shown in Chapter 2. The
incremental energy consumption for the CO2 capture in the cement plant by using the
Ca-looping process could be reduced to as low as 2.3 GJth/ton CO2, which is the
value calculated for the ‘oxy-calciner only’ case at 90% CO2 avoidance in Chapter 2.
However, the energy consumption increases further with a decrease in F0/FCO2 ratio
as well as depending on the deficiency in sorbent performance. Therefore, it can be
concluded that among the CaO-based carbon capture options presented in Chapter 2
and this chapter, the Ca-Cu looping process requires the lowest incremental energy
consumption. If there is an economical way to separate/reuse of CuO/Al2O3 sorbent
from the purge, the energy consumption in the Ca-Cu looping process is possibly
improved further by increasing the F0/FCO2 ratio as demonstrated for the Ca-looping
process in Chapter 2. Moreover, the capital cost requirement of the system can
potentially be reduced by the elimination of an ASU. However, a detailed economic
analysis is necessary to see the impact of the excessive cost of CuO.

3.2 Indirect Calcination Process

In this section, the process integration of the indirect calcination process existing
in the literature (Rodriguez et al., 2011b) into the base cement plant is investigated.
The indirect calcination process allows separation of CO2 from limestone calcination
in a concentrated form. Even though the reference indirect calcination design was
proposed for CO2 capture from a cement plant, its complete integration has not been
analyzed to date. The raw mill, preheater and kiln existing in a conventional cement
plant were not included in the process analysis. Furthermore, only the calcination
reaction was considered while clay decomposition and clinkerization reactions were
not taken into account. The process integration given in this section aims to minimize
the total thermal energy requirement by using excess energy from high temperature
flue gas streams for raw meal preheating as in the conventional cement
manufacturing process. It also considers the use of an additional CO2 capture unit
combined with the indirect calcination process, called hybrid configuration, since the
standalone indirect calcination application can only provide a moderate level of CO2

87
avoidance. For this purpose, an aqueous amine process removes the remaining CO2
relating to clinkerization occurring at the cement kiln and fuel combustion for the
indirect calcination process.

3.2.1 Fundamentals of the Indirect Calcination Process

Figure 3-7 shows the schematic diagram of a novel indirect calcination process
which separates fuel combustion and limestone calcination into two distinct
chambers. The calciner operates at 930ºC for the complete calcination of limestone
in the raw meal. The high temperature solid stream from the combustor (>1000ºC) is
transferred to the calciner to meet the heat requirement for the endothermic
calcination reaction. The combustor temperature is initially set to 1050°C but is later
altered to illustrate the effect of temperature difference on solid flux and solid
circulation rate. The combustor should be designed by adhering to the limits of
temperature and solid flux in the commercial CFB combustion systems. High
temperature CFB combustion systems are well established technologies for ore
roasting, phyrohydrosis of spent potlining and aluminium hydroxide calcination
processes with operating temperatures of 1050ºC, 1200ºC and even up to 1450ºC,
respectively (Reh, 1995), and allowable solid flux rates of between 10-100 kg/m2s
(Bi and Liu, 2012).

CO2 rich
Gas Flue Gas

Heat

Calciner Combustor
(930oC) (1050oC)
High Temperature
Solid Stream

CaO Limestone Air Fuel


(CaCO3)

Figure 3-7 Schematic diagram of the indirect calcination process (Rodriguez et al., 2011b).

88
The high temperatures in the CFB combustor will restrict the fuel characteristics
due to ash softening problems. Therefore, the pet coke used in the kiln of the base
cement plant, is burnt in the combustor. The composition of the pet coke has already
been given in Table 2-5. The use of pet coke with high sulphur content in the
combustor exacerbates CaSO4 formation since it is assumed that all of the SO2
generated in the combustor reacts with CaO and is irreversibly converted to CaSO4.

3.2.2 Process Integration of the Indirect Calcination Process

To prevent the circulation of significant amounts of clay minerals with CaO and
their interaction at the given temperature range, limestone and clay minerals are fed
into two separate raw mills as presented in Figure 3-8. The moisture-free limestone
from the raw mill is fed to a preheater, where its temperature subsequently increases
by heat exchange with hot gas streams from the combustor and cooler. The excess air
from the cooler is initially at 279°C but is further heated up to 880°C in a
regenerative heat exchanger while the CO2-rich gas stream leaving the calciner at
930°C is cooled to 330°C. While the limestone stream from the preheater is sent to
the calciner solely, the product CaO from this reactor is mixed with preheated clay
before the kiln stage.

Since high solid temperatures are achieved in the limestone preheater, the
assumption of 10% calcination in the preheater given in Chapter 2 has been retained.
Accordingly the corresponding CO2 emissions cannot be captured. Having been
cooled to certain temperatures in the limestone preheater, the exit gas stream is able
to cover the heat requirement in the limestone raw mill and fuel drying unit. Since
there is no combustion occurring in the calciner anymore, the tertiary air at 908°C
can be entirely fed to the combustor. The heat requirement in the kiln reduces with
the assumption of complete calcination of limestone in the calciner. All surplus
secondary air at 1025°C, which is not needed in the kiln anymore, can also be
transferred to the combustor. Finally, the excess air from the clinker cooler is able to
cover the remaining requirement in the combustor. In the clay preheater, the flue
gases from the kiln heats up the clay and leaves the system at 110°C.

89
To
Atmosphere

Collected
B/F
Gas Flow Dust
Solid Flow
To Pet Coke F/D
Atmosphere
Primary
Air To
B/F Compressed CO2 Comp B/F
CO2 Atmosphere

Secondary Excess
st nd rd th Air Air
1 2 3 4
Limestone R/M Cal Kiln Cooler
PHE PHE PHE PHE Clinker

Air Air Air


Cooling
To Comb Air
Atmosphere

B/F

1st 2nd 3rd 4th


Clay R/M
PHE PHE PHE PHE

Air Air

Figure 3-8 Schematic diagram of the process integration of a cement plant with an indirect calcination process. Abbreviations: R/M, Raw Mill; B/F, Bag Filter; F/D,
Fuel Drying; PHE, Preheater; Cal, CFB Calciner; Comb, Combustor; CO2 Comp, CO2 Compression.
90
The presence of CaO and SiO2 in a single preheater system leads to formation of
belite (C2S) phase under the conditions of final preheater stage and pre-calciner as
detailed in Chapter 2. Since the CaO and clay minerals are mixed just before the kiln
stage in this configuration, belite formation is postponed. Thus, it would be needed to
adjust the residence time in the kiln by controlling its rotational speed so that the
same clinker composition with the base cement plant can be obtained. Accordingly
an additional reaction step prior to the kiln stage needs to be included in the
simulation of the indirect calcination process where belite formation occurs.

Although air leakages in the raw mills, preheaters and kiln are not so critical for
this configuration and included in its simulation, it is very important to minimize air
leakage in the calciner to prevent the dilution of the CO2 rich stream. Also, part of
the CO2 released in the calciner can be transferred to the combustor (or vice versa) on
the pores and interparticle voids of the circulated solids even though very low gas
transfer rates are expected. In this chapter, air leakages into the new CFB systems
have been neglected as in Chapter 2; however, the negative impacts of air leakages
on CO2 purity can clearly be seen in Chapter 5.

3.2.3 Amine Process and Its Process Integration

The schematic diagram of the amine process that is attached to the base cement
plant is shown in Figure 3-9. To obtain accurate performance predictions for the
absorber and stripper, the add-on amine thermodynamic package in UniSim Design
is employed. Since the primary purpose of this section is to evaluate alternative
carbon capture options rather than optimization of the absorber and stripper designs,
the process conditions given in Ahn et al. (2013) have been adapted. The technical
modelling parameters for the amine process are presented in Table 3-4.

The feed gas stream is initially sent to a cooler, where it is cooled to 32°C and part
of the water is condensed out. It is then pressurized to 1.31 bar by a blower prior to
the absorption stage. The lean amine concentration is set to 30 wt% by adjusting
MEA and water make-up flowrates. The CO2-rich sorbent is pumped to the section at
the bottom of the striper that operates at 1.93 bar. A water wash tower is included for
the recovery of vaporised MEA which is then sent back to the absorber.

91
CO2-depleted Gas

CO2-rich Gas

Water

Absorber MEA, Water Stripper

Flue Gas

Waste Water

Flue Gas Condensate


Lean Solvent Water Wash
Rich Solvent / Reflux Carbon Dioxide
Water /Amine Steam

Figure 3-9 Schematic diagram of the amine process (Ahn et al., 2013).

An important design issue for the amine process is to satisfy the heat duty in
steam stripper for solvent regeneration. The CHP plant configuration reported by the
IEA (2008) is attached to the amine plant configuration, and its schematic diagram is
presented in Figure 3-10. In the CHP plant, the main steam at 500°C and 130 bar is
blown into a turbine for power generation and is then fed to the reboiler at 144.7°C
and 3.5 bar. The power generation in the CHP plant partially/completely fulfils the
power requirements in both base cement and carbon capture plants.

Table 3-4 Main modelling parameters for the amine process.

Unit
Solvent Monoethanolamine (MEA) -
MEA concentration 30 wt%
Absorber feed gas temperature 43.5 °C
Absorber feed gas pressure 1.31 bar
Stripper bottom pressure 1.93 bar
Stripper feed temperature 100 °C
Steam source CHP plant -
CHP plant steam conditions 500 / 130 °C / bar
Steam provided to the reboiler 144.7 / 3.5 °C / bar
Steam after the reboiler 138.5 / 3.5 °C / bar
NOx control Ammonia SCR -
SOx control Wet-limestone FGD -

92
CHP
flue gas
500°C 144.7°C 138.5°C
130 bar Turbine
3.5 bar 3.5 bar

Reboiler
Coal
Air

Boiler 175°C
5 bar

Deaerator
Pump

Condensate
Pump Steam
Lean Solvent

Figure 3-10 Process schematic of the CHP plant (IEA, 2008).

Two different configurations have been assessed for the integration of the amine
process into the base cement plant. The corresponding schematic diagrams shown in
Figures 3-11(a) and 3-11(b) are for the hybrid configuration and the standalone
amine process, respectively. The hybrid configuration is comprised of the indirect
calcination process and an amine process. In this system, the excess air from the
cooler, after being heated up by hot CO2-rich gas from the calciner, is fed to the CHP
plant instead of the limestone preheater. In that way the temperature of the solid
stream in the limestone preheater and relevant partial calcination level can be kept
similar to those in the base cement plant. The CHP plant simultaneously generates
steam for the solvent regeneration and produces electricity. Therefore, the thermal
energy requirement in the CFB combustor located in the hybrid configuration is
expected to be higher compared to that in the standalone application of the indirect
calcination process. The CHP plant configuration employed in this system would be
similar to large scale CFB combustion systems for CHP plants which are
commercially in operation and will be scaled up to 600 MWe in a near future
(Nuortimo, 2013). The conventional CHP plant layout presented in Figure 3-10 has
been modified in the hybrid configuration to include solid circulation between the
reactors while the reference steam conditions have been retained.

93
(a)

Collected
B/F
Gas Flow Dust
Solid Flow
Pet Coke F/D

Primary
To CO2 Air To
B/F B/F
Compression Atmosphere

Secondary Excess
Air Air
1st 2nd 3rd 4th
Limestone R/M SCR Cal Kiln Cooler
PHE PHE PHE PHE Clinker

Air Air Air


Cooling
CHP Air
B/F CO2-depleted
Gas
LP
Steam
1st 2nd 3rd 4th Amine FGD
Clay R/M SCR
PHE PHE PHE PHE

Air Air Compressed CO2 Comp


CO2
94
(b)

CO2-depleted Gas Flow


Gas Solid Flow

LP
Coal Steam Amine CO2 Comp Compressed To
CHP
Air CO2 Atmosphere

Collected
SCR FGD B/F
Dust

Coal F/D To
B/F Pet Coke B/F
Primary Atmosphere
Air
Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M SCR Pre-C Kiln Cooler
Meal PHE PHE PHE PHE Clinker

Air Air Air Air


Cooling
Tertiary Air
Air

Figure 3-11 Schematic diagram of the process integration of a cement plant with (a) hybrid configuration; (b) standalone amine process. Abbreviations: R/M, Raw
Mill; B/F, Bag Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; Cal, CFB Calciner; CHP, Combined heat and power plant; SCR, Selective Catalytic
Reduction Unit; FGD, Flue Gas Desulphurization Unit; Amine, Amine Scrubbing Process; CO2 Comp, CO2 Compression.
95
For the reduction of NOx in the feed gas stream, two SCR units are located
between the raw mills and the preheaters as suggested by IEA (2008). The
temperature of the gas streams at these locations is around 320°C, and the reductant
in the SCR units is selected to be 25% ammonia solution. The flue gas stream from
the bag filters are blown into an FGD unit for SOx removal before it is fed to the
amine process. The solvent flow rate in the amine process is varied in order to
achieve 90% CO2 avoidance in the system.

The second option presented in Figure 3-11(b) shows an end-of-pipe integration


of the amine process in the base cement plant. The same type of coal as in the base
cement plant is burnt in the CHP plant, the composition of which is already given in
Table 2-5. The flue gas stream from the CHP plant, after passing through an SCR
unit, is mixed with the flue gas stream from the base cement plant and is fed to an
FGD unit. The SOx- and NOx-depleted flue gas stream is then transferred to the
amine plant. As before, the CO2 avoidance rate is set to 90% by adjusting the solvent
flowrate in the absorber. The CO2-rich gas stream from the stripper is sent to a CO2
compression unit in both configurations. The design of this unit is identical to the one
detailed in Chapter 2.

3.2.4 Results and Discussion

Figure 3-12 presents the stream properties of the calciner, combustor and kiln
system in the standalone indirect calcination process for the production of 113.0 ton
clinker/h. As pet coke is fed into the combustor instead of coal as in the pre-calciner
of the base cement plant, the ash content of clinker produced in this system is lower.
Thus, the slight reduction in the clinker product rate compared to that of the base
cement plant (113.8 ton clinker/h) can be associated with the reduced flow rate of ash
in the clinker. However, the clinker composition is almost identical in both plants. It
is well-known that the perfect match between clinker compositions facilitates the
calculation of robust mass and energy balances as a result of the enthalpy changes
related to the clinkerization reactions.

96
(15) (14) (13) (12) (11) (10) (9)

(5)
(1) Calciner Combustor Kiln (8)
(4)
(2)
(7)
(3)

(6)

Stream Description Temperature (oC) Mass Flow (ton/h)


(1) CaO/CaCO3 725 129
(2) CaO 930 2048
(3) CaO to kiln 930 78
(4) CaO to combustor 930 1970
(5) CaO to calciner 1050 1972
(6) Clay 750 53
(7) CaO/Clay 845 131
(8) Clinker 60 113
(9) Kiln flue gas 1025 54
(10) Secondary air/Primary air 1025 33
(11) Fuel to kiln 101 3
(12) Combustor flue gases 1050 106
(13) Air mixture 794 100
(14) Fuel to combustor 101 8
(15) CO2-rich gas 930 53

Figure 3-12 Stream properties of the main flows of indirect calcination process integrated into the
base cement plant (Figure 3-8).

The temperature of the CaO/CaCO3 stream (stream 1) entering the calciner is


725°C as a result of heat exchange with the hot gases in the limestone preheater. The
high solid feed temperatures reduce the thermal energy requirement in the calciner so
that it is lower compared to that of the reference indirect calcination configuration
(Rodriguez et al., 2011b) where almost the same quantity of limestone is fed to the
calciner at the ambient temperature. Accordingly the flow rate of the solid recycle is
reduced to 1970 ton/h compared to 3200 ton/h reported in the literature. The recycled
solid stream comprises of mainly CaO (97 wt%) with some ash and CaSO4.

97
40 2800

30 2400

Solid circulation (ton/h)


Solid Flux (kg/ms2)

20 2000

10 1600

0 1200
1020 1040 1060 1080 1100 1120
Combustor Temperature (°C)

Figure 3-13 Variation of solid flux and solid circulation rate in the combustor at different
temperatures.

The temperature of the combustor in the indirect calcination process is altered in a


range of 1025 - 1100°C to evaluate its impact on solid flux and solid circulation rate.
The results are presented in Figure 3-13. The maximum combustor temperature is
limited to 1100°C to prevent the increase in the solid temperature and relevant partial
calcination rate in the preheater. To estimate the solid flux, the superficial velocity of
gases in the reactor can be assumed to be 5 m/s being a typical value for CFB
systems in fast fluidization regime. The solid recycle rate reaches the minimum at
1100°C with a value of 1350 ton/h. The solid flux estimates in the combustor ranges
between 19–37 kg/ms2 at different combustor temperatures and can be handled by the
current status of CFB combustors according to Bi and Liu (2012).

To validate the assumption of complete calcination in the calciner, a further


analysis has been conducted. Martinez et al. (2010) claimed that the calcination
reaction is chemically controlled and the internal mass transfer resistance is
negligible up to a particle size of 300 μm.

98
They proposed the following equation based on the reaction kinetics to calculate
the time required for achieving full calcination (tc*):

3𝑋𝑎𝑣𝑒
𝑡𝑐∗ = (Eq. 3.2)
𝑘𝑐 (𝐶𝐶𝑂2 ,𝑒𝑞 −𝐶𝐶𝑂2 )

where Xave (mol CaCO3/mol Ca) is the CaCO3 content of the solid stream entering
the calciner, kc (m2kmol−1s−1) is the kinetic constant for the calcination reaction.
CCO2,eq (kmol/m3) and CCO2 (kmol/m3) are the CO2 concentration at equilibrium and
in the gas phase respectively. The value of kinetic constant is estimated by the
following equation:

−𝐸𝑎𝑐
𝑘𝑐 = 𝑘𝑐0 exp � � (Eq. 3.3)
𝑅𝑇

where kc0 (m3kmol−1s−1) is the pre-exponential factor, and Eac (kJ/mol) is the
activation energy of calcination. The kinetic parameters of “limestone A” given in
the reference have been selected for use in the calculations, for which kc0 is 2.1×106
m3kmol−1s−1, and Eac is 112.4 kJ/mol (Martinez et al., 2010).

The Xave value in Eq. 3.2 is equal to 1 because only fresh limestone is fed to the
calciner. The required calcination times for complete calcination at different calciner
temperatures are presented in Figure 3-14. Two different calcination atmospheres
have been examined: (i) pure CO2 and (ii) 95% CO2. The latter is included to
consider any unhindered air leakages. It can be proven that the calcination
temperature of 930°C selected in this study would allow reaching complete
calcination rapidly (less than 25 seconds) even under pure CO2 atmosphere.
Although a detailed design of the calciner has not been investigated, it is expected
that estimated solid residence times can easily be achievable in a BFB calciner,
which is a potential candidate for the application of this system. At lower
temperatures, however, the required calcination time for complete calcination
increases further, i.e. over 3 minutes at 900°C.

99
940
100% CO₂
95% CO₂

930

Calcination Temperature ( C)
920

910

900

890
0 40 80 120 160 200
Calcination Time (s)

Figure 3-14 Required calcination time at different temperatures and calcination atmospheres to
achieve complete calcination.

The process performances of the proposed schemes in this section are compared
in Table 3-5. The total thermal energy requirement in the standalone indirect
calcination process is slightly lower than that of the base cement plant (see Table 3-
3) which can be linked to higher level of heat recovery from the excess air stream.
The main energy penalty for this system results from the CO2 compression unit. The
indirect calcination process provides approximately 56% CO2 avoidance without a
need of any additional carbon capture technology. By incorporating the amine
process into the indirect calcination process, 90% CO2 avoidance can be achieved.
Although the thermal energy requirement of the cement plant in the hybrid
configuration seems to be much lower than the other options, it is because of the pre-
calciner duty that is included in the heat requirement of the CHP plant.

In the hybrid configuration, the concentration of the CO2 in the feed gas entering
the absorber is around 14 mol% which is a typical value for a coal-fired combustor
whereas this value is around 20 mol% for the standalone application of the amine
process. The difference is due to the additional CO2 emissions resulting from the
calcination of limestone. Besides, the molar flow rate of the feed gas is less in the
hybrid configuration. In order to reach the avoidance rate of 90%, the CO2 capture
efficiency is set to 85% in the hybrid configuration, while up to 95% capture
efficiency is required in the standalone amine process because of the excessive CO2

100
generation in the CHP plant. For this reason, the duty of reboiler increases to 3.96
MJth/kg CO2 in the latter.

The power requirement of the capture plant in the hybrid configuration is


calculated to be 19.7 MWe. This requirement arises from the CO2 compression unit
(11.7 MWe), pumps and compressors in the amine process as well as FGD and SCR
units. The total electricity requirement of 42.4 MWe in the standalone amine process
can be met by generation in the CHP plant where an extra 7.8 MWe can potentially
be exported or sold to the grid. Nevertheless, the gross output of the CHP plant
located in the hybrid configuration is not enough to meet the entire power
requirement so remaining 19.8 MWe should be imported. Considering all thermal
energy and power requirements, the incremental energy consumption goes up to 8.2
GJth/ton CO2 avoided in the standalone amine process while it is calculated to be 3.3
GJth/ton CO2 avoided for the hybrid configuration at the same CO2 avoidance rate.
The incremental energy consumption in the standalone indirect calcination process is
found to be only 0.9 GJth/ton CO2 avoided, which reflects the power requirement in
the CO2 compression unit.

101
Table 3-5 Comparison of the standalone indirect calcination and amine processes, and the hybrid configuration.

Standalone Standalone Amine Hybrid


Indirect Calcination Process Configuration
Thermal energy requirement (GJth/ton clinker)
in the cement plant 3.08 3.13 0.76
in the CHP plant - 7.70 4.33
Amine process
Feed gas composition (mol %) - 2% H2O, 4% O2, 2% H2O, 7% O2,
74% N2, 20% CO2 77% N2, 14% CO2
CO2 in feed (kmol/h) - 3891.2 1425.0
CO2 capture efficiency (%) - 95 85
CO2 lean loading (mol/mol) - 0.23 0.23
CO2 rich loading (mol/mol) - 0.46 0.51
Solvent regeneration energy requirement (MJth/kg CO2) - 3.96 3.28
Power consumption (MWe)
Cement plant auxiliaries 13.7 13.7 13.7
Capture plant
CO2 compression 5.7 15.0 11.7
Amine plant auxiliaries - 13.7 8.0
Power generation (MWe) - 50.2 13.6
Net power (MWe) −19.4 7.8 −19.8
CO2 intensity (ton CO2/ton clinker) 0.36 0.08 0.08
CO2 avoidance rate (%) 55.6 90.0 90.0
Incremental energy consumption (GJth/ton CO2 avoided) 0.9 8.2 3.3
102
3.3 Membrane Separation Process

The other advanced technology that has been mostly studied for the purpose of
carbon capture from the power generation industry is membrane separation (Zhao et
al., 2010; Merkel et al., 2012; Ramasubramanian et al., 2012; Zhai and Rubin, 2013).
Even though this technology is still under development and has only been applied to
small scale natural gas sweetening and oxygen enrichment applications to date (He
and Hagg, 2012), it has been identified as a very promising option for the removal of
CO2 from cement flue gases (ECRA, 2007). However, its sensitivity to sulphur
compounds and other trace elements in addition to difficulties in capture of 90% CO2
in flue gas with a high-purity permeate stream in a single membrane stage has been
pointed out as unfavourable characteristics. In the case of the latter, the use of
multiple stages and/or recycling has often been suggested as a solution. Subsequent
studies have focused on improving the properties of membranes such as CO2
permeance and CO2/N2 selectivity (Zhao et al., 2010). Merkel et al. (2012)
successfully developed a multi-stage analysis focusing on a retrofit integration and
recycling of part of the CO2 to the boiler of a power plant. A PolarisTM membrane
with a CO2 permeance of above 1000 GPU (where 1 GPU is equal to 10−6
cm3(STP)/cm2·s·cmHg or a CO2 permeability of 1000 Barrers assuming a 1-μm-
thick membrane) and a CO2/N2 selectivity of 50 at 30°C has been used in their
analysis.

The objective of this section is to give an overall insight into the application of the
membrane process in the base cement plant for carbon capture. An analysis of a
variety of multi-stage membrane configurations has been carried out. The membrane
unit operation developed as an extension in UniSim Design by the Carbon Capture
Group at University of Edinburgh (Bocciardo et al., 2012; Bocciardo et al. 2013) has
been employed in this section.

3.3.1 Modelling of the Membrane Process

The modelling of a membrane for CO2 separation is not in the scope of this
section. In Bocciardo (2014), different membrane models have been developed that
allow predicting the behaviour of industrially available membrane permeators. These

103
models have been incorporated into the base cement plant simulation to avoid 90%
CO2 with the target purity of ≥95 mol%. Under the assumptions of: (i) no pressure
drop, (ii) ideal and isothermal behaviour, and (iii) constant permeance along the
module, the models chosen for the process simulations are:

• Non-disperse plug-flow: the set of differential equations which represent the


mass balance for a counter-current flow pattern using a sweep gas in the
permeate are for retentate side (Eq. 3.4) and permeate side (Eq. 3.5).

𝑑𝐹𝑟 𝑖=𝑁𝐶
= − ∑𝑖=1 𝜋𝑖 (𝑃𝑟 𝑥𝑖𝑟 − 𝑃𝑝 𝑥𝑖𝑝 ) (Eq. 3.4)
𝑑𝐴

𝑑𝐹𝑝 𝑖=𝑁𝐶
= ∑𝑖=1 𝜋𝑖 (𝑃𝑟 𝑥𝑖𝑟 − 𝑃𝑝 𝑥𝑖𝑝 ) (Eq. 3.5)
𝑑𝐴

where F is the molar flow rate, P is the pressure, π is the permeance, and x is the
molar fraction. Besides, A is the area, and NC refers to number of components. The
suffixes r and p refer to retentate and permeate sides, respectively. Following the
previous study on post-combustion capture from coal-fired power plants (Bocciardo
et al., 2013) where counter-current with sweep has been marked as the best option
for high-recovery stages, part of the retentate is chosen as sweep and is recycled to
the permeate. The use of a sweep stream reduces the CO2 partial pressure in the
permeate and thereby improves the driving force of separation. In this way, the same
overall recovery can be achieved by keeping relatively small membrane areas. This
flow pattern can be found in hollow-fibre modules, where fibres with diameters up to
mm scale are assembled to achieve area/volume ratio of 10000 m2/m3 (Baker, 2004).

• 2-D Cross-flow: a cross-flow formulation has been adapted to model separation


through spiral-wound permeators. The model equations are presented as follows
for retentate side (Eq. 3.6) and permeate side (Eq. 3.7), where s and v are the
retentate molar flow rates per unit of width (w) and permeate molar flow rate per
unit length (l), respectively.

𝑑𝑠 𝑖=𝑁𝐶
= −2 ∑𝑖=1 𝜋𝑖 (𝑃𝑟 𝑥𝑖𝑟 − 𝑃𝑝 𝑥𝑖𝑝 ) (Eq. 3.6)
𝑑𝑙

𝑑𝑣 𝑖=𝑁𝐶
= 2 ∑𝑖=1 𝜋𝑖 (𝑃𝑟 𝑥𝑖𝑟 − 𝑃𝑝 𝑥𝑖𝑝 ) (Eq. 3.7)
𝑑𝑤

104
3.3.2 Selection of an Optimal Feed Gas Stream

Two feed gas locations have been identified in this section for the integration of a
membrane separation into the base cement plant. The flue gas stream leaving the bag
filters can be fed to the membrane separation process as an end-of-pipe integration.
Alternatively, the flue gas stream from the preheater stage, which has a higher CO2
concentration, can be preferable for CO2 capture. The corresponding schematic
diagrams for these options are presented in Figure 3-15 and Figure 3-16,
respectively.

For the end-of-pipe integration shown in Figure 3-15, there is no significant


modification required in the base cement plant except for the additional FGD and
SCR units. The NOx and SOx emission levels can reach up to 1450 ppm and 1200
ppm, respectively in a cement process (IEA, 2008). The negative effects of minor gas
components including NOx and SOx on membrane separation performance have been
reported in Scholes et al. (2011), but further studies are still required to determine
acceptable limits for such impurities. Since majority of the experimental studies have
been carried out without having NOx and SOx in the feed gas stream, SCR and FGD
units have been included in the process simulations as in the reference studies
(IECM, 2012; NETL, 2012). The SCR unit is installed between the preheater and
raw mill, and the FGD unit is located after the bag filters as in the previous section.

In the alternative option presented in Figure 3-16, the flue gas stream at 320°C
from the preheater stage is first passed through an SCR unit and then an electrostatic
precipitator (ESP) where solid particulates in the gas stream are removed. Since the
flue gas stream leaving the preheater is directly sent to the capture unit rather than
the raw mill, the heat requirement in this unit and fuel drying system cannot be
satisfied anymore. However, it would be possible to transfer heat from the gas stream
leaving the ESP to the CO2-depleted gas stream from the membrane process in a
regenerative heat exchanger. The preheated CO2-depleted gas stream in addition to
the excess air from the cooler will be sufficient to meet the heat demand in the raw
mill. Finally, the feed gas stream after being cooled in the regenerative heat
exchanger flows through an FGD unit prior to the membrane process.

105
CO2 Compressed
Comp CO2

CO2- depleted Gas Flow


Membrane Gas Solid Flow

Collected
FGD B/F
Dust

Coal To
B/F Pet Coke F/D Primary B/F
Atmosphere
Air
Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M SCR Pre-C Kiln Cooler
Meal PHE PHE PHE PHE Clinker

Air Air Air Air Cooling


Air
Tertiary Air

Figure 3-15 Schematic diagram of a cement plant with a membrane-based CO2 capture unit (The end-of-pipe option). Abbreviations: R/M, Raw Mill; B/F, Bag Filter;
F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; SCR, Selective Catalytic Reduction Unit; FGD; Flue Gas Desulphurization Unit; CO2 Comp, CO2
Compression.
106
CO2 Compressed
Comp CO2

Membrane CO2-depleted
Gas
To
Atmosphere
FGD
Gas Flow
To Collected
Solid Flow B/F
Atmosphere Dust
HE To
Atmosphere

B/F Coal F/D B/F


ESP Pet Coke Primary
Air
Secondary Excess
st nd rd th Air Air
Raw 1 2 3 4
R/M SCR Pre-C Kiln Cooler
Meal PHE PHE PHE PHE Clinker

Air Air Air Air Cooling


Tertiary Air Air

Figure 3-16 Schematic diagram of a cement plant with a membrane-based CO2 capture unit (An alternative integration option). Abbreviations: R/M, Raw Mill; B/F,
Bag Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; SCR, Selective Catalytic Reduction Unit; FGD; Flue Gas Desulphurization Unit; CO2 Comp, CO2
107

Compression; HE, Heat Exchanger.


Table 3-6 Composition of the flue gas stream after the bag filters (end-of-pipe) and the preheater.

End-of-pipe After Preheater


(mol %) (mol %)
CO2 22.4 32.3
N2 58.1 59.7
O2 6.2 2.4
H2O 13.2 5.5
SO2 0.1 0.1

Molar flow (kmol/h) 9398.0 6522.0

As can be seen in Table 3-6, the major benefits of the alternative integration
where the preheater gas stream is selected as a feed gas are greater CO2
concentration (32.3 mol% vs. 22.4 mol%) and lower volumetric gas flow as a result
of eliminating the air leakages and moisture in the raw mill.

3.3.3 Membrane Process Configurations

For the membrane separation process, the major concern is dealing with high
volume, low pressure feed gas streams when it is implemented in industrial scale
applications. The CO2 permeation of gas separation membranes is directly related to
partial pressure difference between the feed and permeate sides. Thus, it is important
to select the optimal pressure ratio as this also directly influences the performance of
membranes as their properties and CO2 concentration in the feed gas do.

The cost and energy requirements are the key concerns for generating a pressure
difference across a membrane. Pressure ratios between 5 and 10 have been
determined as economically affordable for this application (Merkel et al., 2010).
Very high pressure ratios can be achievable by the utilization of compressors on the
feed gas or by drawing a vacuum on the permeate side, but the main criterion is the
selection of an economical way. There are three major advantages of using a gas
compressor: (i) the capital cost of a gas compressor is half of that for a vacuum pump
(IECM, 2012), (ii) the use of a compressor requires less membrane area compared to
a vacuum (Merkel et al., 2010), and (iii) it would be possible to recover part of the
energy spent for compression by employing a turbo-expander (Merkel et al., 2010).
However, the vacuum operation is favoured owing to reduced energy consumption.
A vacuum pressure of 0.22 bar slightly higher than the lowest practical vacuum

108
pressure (0.2 bar) for CO2 capture applications is selected in this study following the
reference (Merkel et al., 2010). Hence, the feed stream is initially compressed to 1.1
bar to have a pressure ratio of 5.

Promising results have been reported for improvements in membrane properties


(Merkel et al., 2010; Favre, 2011). In addition to enhanced membrane permeability
and selectivity, researchers have been focusing on the development of highly stable
membranes that would be suitable for industrial scale applications. Membranes with
higher selectivities are required to make the CO2 separation process more attractive.
In this study, the commercial membrane, PolarisTM (Merkel et al., 2012) with a CO2
permeance of 1000 GPU and CO2/N2 selectivity of 50 has been used adhering to the
latest developments in this field. CO2/O2 and CO2/H2O selectivities have been
estimated from the data available in NETL (2012) with the values of 10 and 0.2,
respectively. A feed gas temperature is set to 30°C following the reference (Merkel et
al., 2010), but the dependence of permeability on the gas temperature is not taken
into account. It is not possible to capture 90% of the CO2 at very high CO2 purities
from a typical feed gas stream by using a single membrane stage regardless of the
type of membrane. Therefore, multi-stage configurations are necessary to reach
higher degrees of separation as well as the purity.

Two membrane models, 2D cross-flow and counter-current with sweep, called


‘cross-flow’ and ‘counter-current’ respectively from now on, have been used in the
multi-stage analysis. Four different dual-stage configurations have been evaluated as
illustrated in Figure 3-17. The flue gas stream leaving the preheater stage (see Table
3-6) has been selected as a feed gas in all configurations. The feed gas stream is
initially compressed to 1.1 bar and then cooled to 30°C while the condensed water is
separated. In the first configuration (Conf. 1) shown in Figure 3-17(a), the counter-
current stage has been located in the first stage as it provides high recovery, and
followed by the cross-flow stage for high purity in the second stage. The feed gas
stream is mixed with recycled retentate from the cross-flow stage and then fed to the
counter-current stage. Part of the retentate stream, that is 10% for a preliminary
analysis, from the counter-current stage is recycled back to the permeate side as a
sweep. The remaining retentate stream is marked as CO2-depleted gas stream.

109
(a)

Counter-current
Sweep
Cross-flow
Flue gases from Expander
the FGD unit
Compressor CO2 - rich
Gas
H2O Vacuum Pump Compressor Vacuum Pump

CO2-depleted H 2O
Gas

(b)
CO2-depleted
Gas

Expander
Cross-flow
Counter-current
Module
Sweep

CO2 - rich
Flue gases from Gas
the FGD unit
Vacuum Pump
Compressor

H 2O
Compressor Vacuum Pump

H 2O

(c)

Counter-current Expander
Sweep
Counter-current
Flue gases from Sweep
Expander
the FGD unit
Compressor
CO2 - rich
H 2O Vacuum Pump Compressor Gas
Vacuum Pump
CO2-depleted H 2O
Gas

(d)
CO2-depleted
Gas

Expander

Counter-current
Counter-current
Sweep
Sweep

Flue gases from Expander


the FGD unit CO2 - rich
Compressor Gas
Vacuum Pump
H2O

Compressor Vacuum Pump

H2O

Figure 3-17 Schematic diagrams of the proposed membrane process configurations; (a) Conf. 1, (b)
Conf. 2, (c) Conf. 3, (d) Conf. 4.

110
The permeate stream from the counter-current stage is compressed to 1.1 bar,
cooled to 30°C and sent to the cross flow stage. The CO2-rich gas stream from the
cross-flow stage is ready to be transferred to the CO2 compression unit, which has a
similar configuration to the one detailed in Chapter 2. In addition, three other
alternative dual stage configurations have been examined. In Conf. 2, presented in
Figure 3-17(b), a cross flow stage is followed by a counter current stage. A dual
counter current configuration, Conf. 3, is presented in Figure 3-17(c). Finally,
another dual counter current design with different gas recycle option is given in
Conf. 4 shown in Figure 3-17(d). For all cases, the feed gas stream is taken from the
preheater and the sweep/retentate (S/R) ratio is fixed at 0.1.

3.3.4 Results and Discussion

The outcomes of the dual-stage membrane configurations providing 90% CO2


avoidance are compared in Table 3-7. It should be noted that the tabulated results are
the best estimates obtained by multiple trials to achieve the target avoidance rate and
purity with minimum energy consumption and membrane area. However, an
optimization work would allow further reduction in the current estimates. The power
consumptions of the pumps and compressors, including those in the CO2
compression unit, are converted to thermal energy by using a power plant efficiency
of 0.4 as before. It is clear from the results that the target purity can be achieved
successfully by only Conf. 1 and Conf. 2, and among these two configurations,
Conf.1 requires slightly less membrane area and energy consumption. On the other
hand, the proposed dual counter-current configurations (Conf. 3 and Conf. 4) could
reach up to 92.5 mol% CO2 purity. Nevertheless, they are not favourable as Confs. 1
and 2 since an additional step is needed to increase the CO2 purity further up to 95
mol%, which definitely increases the energy consumption and/or membrane area as
well as the capital cost requirement. By comparing the energy consumptions and
membrane area requirements for different configurations, it can be proven that
further investigations should be based on Conf. 1 as it would potentially provide the
lowest cost requirement for the carbon capture in the current analysis.

111
Table 3-7 Comparison of various membrane configurations at 90% overall CO2 avoidance rate.

Total Incremental
membrane area energy consumption CO2 purity
Configuration (×103 m2) (GJth/ton CO2 avoided) (mol %)
After preheater
Conf. 1 271 2.0 >95.0
Conf. 2 272 2.2 >95.0
Conf. 3 280 2.0 92.5
Conf. 4 172 1.9 70.0
End-of-pipe
Conf. 1 402 2.1 >95.0

To determine the impact of different feed gas conditions, Conf. 1 is also employed
in the end-of-pipe integration presented in Figure 3-15. The higher volumetric gas
flow rate and lower CO2 concentration of the feed gas stream in this option results
approximately 48% increase in the membrane area. In addition, the energy
consumption increases to 2.1 GJth/ton CO2 avoided. The results indicate that the
major difference in terms of process economics, when two different ways of
integration are compared, will potentially be the cost of membrane modules since the
power requirements are very similar in both options.

3.3.4.1 Sensitivity Analysis of Membrane Parameters

It is necessary to investigate the impact of S/R ratio fixed initially at 0.1 on


membrane area and energy consumption estimations. Furthermore, a sensitivity
analysis on the CO2 permeance and CO2/N2 selectivity will reveal how the recent
developments on membranes properties will influence the current estimates. To
conduct a sensitivity analysis, the feed gas conditions given in Table 3-6 for the
preheater gas stream and Conf. 1 have been used. The variation in the membrane
area and energy consumption between 0-0.3 S/R is presented in Figure 3-18. It is
observed that the membrane area increases significantly when S/R reduces. This
proves that sweep stream improves the driving force of separation. On the other
hand, very high sweep flow implies higher duty for the vacuum pump at the
permeate side. Thus, following the minimum energy consumption at 0.075 S/R, a
sharp increase is encountered after 0.1 S/R. Since none of examined S/R values
provides minimum membrane area and energy consumption simultaneously, an
economic analysis is needed to reveal the best value for S/R.

112
500 2.3

Incremental Energy Consumption


400
2.2
Total Membrane Area

(GJth/ton CO2 avoided)


(×103) [m2]
300
2.1
200

2
100

0 1.9
0 0.1 0.2 0.3 0.4
Sweep/Retentate (S/R) Ratio

Figure 3-18 The impact of sweep/retentate ratio (S/R) on membrane area and incremental energy
consumption.

In the second part of the sensitivity analysis, the CO2 permeance assumed to be
1000 GPU initially has been varied between 1000-12000 GPU by keeping the
CO2/N2, CO2/O2 and CO2/H2O selectivities constant, and the results are presented in
Figure 3-19(a). No change in the energy consumption estimates is observed while the
major impact of the CO2 permeance appears to be an increase in the membrane area.
Membranes having very low CO2 permeance require very high membrane areas, and
it is shown here that the membrane area can be potentially reduced to 23×103 m2
when the CO2 permeance is increased to 12000 GPU. The impact of CO2/N2
selectivity on both membrane area and consumption is significant as shown in Figure
3-19(b). The energy consumption estimates reduce gradually with an increase in the
CO2/N2 selectivity and reach the maximum at the lowest selectivity. By changing the
CO2/N2 selectivity from 25 to 75, a variation of 12% in the energy consumption
estimates is observed. Clearly the duty of the vacuum pump makes the largest
contribution at low CO2/N2 selectivities. An opposite impact has been observed on
membrane area estimates which show increase at high selectivities because of the
reduction in the driving force for CO2 permeation. The membrane area increases to
307×103 m2 at the selectivity of 75.

113
(a)

300

Total membrane area

200
(×103) [m2]

100

0
0 4000 8000 12000 16000
CO2 Permeance (GPU)

(b)

400 2.2

Incremental Energy Consumption


(GJth/ton CO2 avoided)
Total membrane area

300
2.1
(×103) [m2]

200

2
100

0 1.9
20 40 60 80
CO2/N2 Selectivity

Figure 3-19 The effects of (a) the CO2 permeance on membrane area; (b) CO2/N2 selectivity on
membrane area and incremental energy consumption.

114
The volume of each membrane module can roughly be estimated by using the
membrane area and packing density data given in Bocciardo (2014), where an
average packing density was reported to be 1300 m2/m3 for the cross-flow spiral-
wound modules and 7300 m2/m3 for the counter-current modules. Therefore, the total
membrane area of 307×103 m2 corresponds to an approximate volume of 60 m3.

3.4 Concluding Remarks

This chapter describes the application of alternative carbon capture technologies


to reduce the CO2 emissions from the base cement plant. In the first section, the Ca-
looping process has been coupled with a CLC cycle to eliminate the requirement of
an energy intensive ASU. Within the different ways of solid transfer between the
reactors, a process configuration with a solid route of carbonator → calciner → air
reactor → carbonator has been selected to enable heat transfer between the air reactor
and the calciner as this is a requirement to avoid unfeasible heat duties in the capture
system. As a method of heat transfer, a hot solid stream from the air reactor has been
recycled to the calciner operating at a lower temperature. The air reactor operates at
950°C to provide a sufficient temperature difference. The upper limit of 0.15 F0/FCO2
has been determined to restrain the amount of CuO to a value of 1 wt% in the cement
raw meal. To achieve 90% CO2 avoidance, the capture efficiency of the carbonator is
varied between 88 - 90%. The minimum incremental energy consumption for this
process and for the others estimated in Chapters 2 and 3 are reported in Table 3-8 for
ease of discussion. It was indicated that the energy consumption of the Ca-Cu
looping process can be fairly lower than that of the Ca-looping process depending on
F0/FCO2 ratio and sorbent performance.

The indirect calcination process can provide partial CO2 reduction and energetic
efficiency improvements when it is properly located in the cement plant.
Approximately 56% of CO2 emissions from the cement plant can be prevented
without requirement of an additional post-combustion carbon capture process. An
aqueous amine process was coupled with the indirect calcination process to increase
CO2 avoidance rate further up to 90%. The incremental energy consumption of the
hybrid configuration was found to be lower than that of the standalone integration of

115
the amine process (3.3 GJth/ton CO2 avoided vs. 8.2 GJth/ton CO2 avoided), but it
was almost double of that of the Ca-Cu looping process.

Finally, a potential strategy for effective process integration of a membrane


separation process into the base cement plant has been proposed. Two feed gas
locations have been identified, among which the preheater gas stream requires less
energy consumption and membrane area due to higher CO2 concentration and lower
volumetric flow rate. Since it is not possible to reach very high capture rates along
with stringent CO2 purity requirements by a single stage membrane, a membrane
configuration comprised of a counter-current module with a sweep following by a
cross-flow module has been selected as an optimal configuration as it provides the
target CO2 avoidance rate and purity at minimum expense. It was concluded by
comparing the energy consumption estimates that membrane separation can be a
promising alternative for the Ca-looping process, but an economic study is required
to support this argument.

Table 3-8 Minimum incremental energy consumption estimates for the carbon capture processes
evaluated in Chapters 2 and 3.

Incremental Energy Consumption


(GJth/ton CO2 avoided)
Ca-looping process 2.5
Ca-Cu looping process 1.7
Indirect calcination process 0.9
Hybrid configuration 3.3
Standalone amine process 8.2
Membrane process 2.0
‘Oxy-calciner only’ case 2.3

116
Chapter 4 Economic Analysis

The final comparison of different CO2 capture technologies should take economic
criteria into account. Thus, the aim of this chapter is to provide a simple and
transparent economic analysis for the estimation of costs of cement production and
CO2 avoided associated with the capture technologies evaluated in Chapters 2 and 3.
Although the literature regarding carbon capture economics in the cement industry is
limited, the available methodologies are also summarized in this chapter.

4.1 A Review of Existing Studies

For the indirect calcination process, Rodriguez et al. (2008b) used the following
equation for the calculation of cost of CO2 avoided (AC).

∆𝑇𝐶𝑅 𝐹𝐶𝐹
+𝐹𝐶 ∆𝐻+𝐶𝑂𝐸𝑐𝑜𝑚𝑝
𝐴𝐶 = 𝐶𝐹
−1 −1 (Eq. 4.1)
𝐶𝑂2 𝑡 𝑐𝑒𝑚𝑒𝑛𝑡𝑟𝑒𝑓 −𝐶𝑂2 𝑡 𝑐𝑒𝑚𝑒𝑛𝑡𝑐𝑎𝑝𝑡𝑢𝑟𝑒

where TCR is the specific total capital requirement; FCF is the fixed charge factor
(to amortize the TCR for the plant lifetime and discount rate assumptions); CF is the
capacity factor; FC and COEcomp refer to the fuel and compressor electricity costs,
respectively; and ΔH is the thermal energy requirement. The terms in the
denominator are the CO2 intensities of a reference cement plant with and without
carbon capture. FCF and CF values are fixed at 0.1 and 0.9, respectively. The
assumption implicit in Eq. 4.1 is that the capture plant is of a type and design
equivalent to the reference cement plant. Thus, when a capture unit is incorporated,
Eq. 4.1 calculates the increment in the total cost and allows for comparison of the
average cost of reducing CO2 emissions for the same product capacity base plant.
The TCR of the capture plant is critical to calculations using Eq. 4.1, and in the
absence of a detailed cost analysis for the indirect calcination process, its similarity
of existing CFB combustion systems is often taken into account. In Rodriguez et al.
(2008b), the cost of the CFB combustor was assumed to be 30% of the TCR of an
equivalent CFB power plant, while the compression unit cost was estimated by using
86 $/kW, which was given by Rubin et al. (2007).

117
The latest article on the indirect calcination process (Rodriguez et al., 2011)
proposes a more detailed economic analysis. It also incorporates the cost of a steam
cycle because the excess heat recovered from the high temperature gas streams was
used to drive a steam cycle. The reference cost of 500 $/kWe given in the DOE report
(2007) was employed for the estimation of the steam cycle cost. Since all excess heat
was used for power generation instead of reducing the thermal energy requirement in
the calciner by preheating the solid feed, as proposed in Chapter 3, there was surplus
power generation. A revenue of 0.05 $/kWh was included, under the assumption that
the surplus electricity can be sold to the grid, which gave a negative electricity cost
estimation. The avoided cost was calculated to be 12 $/ton CO2 at 38% efficiency.
The primary reason behind such a low avoided cost is the profit gained from surplus
power generation.

The economic analysis proposed by Romeo et al. (2010) was applied to a system
comprised of a power plant, a cement plant and a Ca-looping carbon capture process.
Romeo et al. (2012) also employed Eq. 4.1 in their study to estimate the cost of CO2
avoided. The specific TCR of a reference air-fired power plant was assessed to be
1100 €/kWe, while that of the calciner was estimated by assuming that the cost of the
boiler represents 40% of the TCR of the power plant. Even though the need to use
scale factor in the cost estimations was mentioned because the size of the reference
plant differs from the actual capacity, this option was skipped in the calculations.
Unfortunately, there is not any reference that particularly presents a way to estimate
the cost of a carbonator in the literature. Hence, to estimate the cost of the carbonator
in their study, Romeo et al. (2010) calculated the TCR of an air-fired power plant
based on the net power generation capacity of the capture plant, assuming that the
carbonator and heat exchanger costs can replace the cost of the boiler in the power
plant, including the steam cycle cost. ASU and compression costs were added
according to the methodology of Rubin et al. (2007). The cost of CO2 avoided was
reported to be 12.4 €/ton CO2 for this system, which was stated to be lower than the
individual application of the Ca-looping process for power and cement plants.

118
The process scheme presented in Rodriguez et al. (2012) has some similarities
with the system explained in Chapter 2, except for the way of integrating the capture
plant. In the reference, the economic performances of the oxy-fired CFB calciner and
Ca-looping cases were compared. For the oxy-fired calciner case, the cost of a pre-
calciner in the base cement plant was discounted since this unit was replaced by an
oxy-fired calciner. A cost of 20 $/annual ton clinker was assumed for the existing
pre-calciner. Since the economic data for an oxy-fired calciner are scarce, those
available for an equivalent oxy-fired CFB power plant were taken from the European
Commission’s SETIS Energy Information System (SETIS, 2013). The calculated
TCR was 3125 $/kWe which covers the costs of an oxy-fired boiler, an ASU, a gas
processing system including CO2 compression, a steam cycle and utilities. For the
Ca-looping case, the carbonator cost was calculated as a modest fraction (10%) of
the TCR of the additional oxy-fired CFB power plant. The cost of cement (COCref)
was given according to the following equation:

∆𝑇𝐶𝑅𝑟𝑒𝑓 𝐹𝐶𝐹
𝐶𝑂𝐶𝑟𝑒𝑓 = + 𝐹𝐶 𝐻𝑟𝑒𝑓 + 𝐶𝑂𝐸 (Eq. 4.2)
𝐶𝐹

This is very similar to Eq. 4.1 but was adapted for cement production. Here,
TCRref refers to the total capital cost requirement of a reference cement plant, and
Href is the thermal energy requirement. The cost of electricity (COE) covers the entire
electricity requirements for the plant. For the proposed carbon capture processes,
TCR and H values increase because of the additional reactors, steam cycle, ASU, and
gas processing system, but the COE can be reduced, since majority of the power
requirement can potentially be satisfied by a heat recovery steam cycle. The cost of
CO2 avoided was then calculated according to the following equation:

𝐶𝑂𝐶𝐶𝑂2−𝑐𝑎𝑝𝑡𝑢𝑟𝑒 − 𝐶𝑂𝐶𝑟𝑒𝑓
𝐴𝐶 = (Eq. 4.3)
(𝐶𝑂2 ⁄𝑡𝑜𝑛 𝑜𝑓 𝑐𝑒𝑚𝑒𝑛𝑡)𝑟𝑒𝑓 −(𝐶𝑂2 ⁄𝑡𝑜𝑛 𝑜𝑓 𝑐𝑒𝑚𝑒𝑛𝑡)𝐶𝑂2−𝑐𝑎𝑝𝑡𝑢𝑟𝑒

where ‘CO2-capture’ and ‘ref’ refer to the capture and reference cases, respectively,
and the CO2 intensity estimations in the denominator reflect the process simulations.
With these assumptions, the cost of CO2 avoided was calculated to be 16 $/ ton of
CO2 and 23 $/ton of CO2 for the oxy-fired calciner and Ca-looping systems,
respectively.

119
In a more recent paper (Romano et al, 2013), the combined cost of a calciner and
a carbonator was estimated using the following equation as a function of reactor
volume (V) and thermal energy requirement in the calciner (Q).

𝑄𝐿𝐻𝑉,𝑐𝑎𝑙𝑐 𝑆𝐹,𝑄 𝑉𝑐𝑎𝑙𝑐 𝑆𝐹,𝑉 𝑉𝑐𝑎𝑟𝑏 𝑆𝐹,𝑉


𝐶𝐶𝑎𝐿 [€] = 𝐶0 �𝛼 � � + (1 − 𝛼) � � + (1 − 𝛼) � � �(Eq. 4.4)
𝑄0 𝑉0 𝑉0

where scale factor values, SF,Q and SF,V, are assumed to be 0.9 and 0.67,
respectively, and α is assumed to be 0.85, representing the ratio of heat transfer
surfaces on the total cost of a cooled CFB reactor. Additionally, parameters C0, Q0
and V0 refer to the TCR of the boiler, the thermal energy input, and the volume of the
boiler, respectively.

Liang and Li (2012) conducted a more detailed economic analysis based on a case
study of a cement plant in China. An amine-scrubbing post combustion process was
implemented to capture 85% of the CO2 from a cement plant. Through a static cash
flow analysis, the cost of CO2 avoided by the proposed design, including a new 200
MWe CHP plant, was estimated to be 70 $/ton CO2 at a 14% discount rate, with the
assumption of a lifetime of 25 years. Although the TCR of the system was calculated
by estimating the cost of each unit, the details of the calculation method were not
exhibited. Li et al. (2013) added an oxy-fired system for comparison with the amine-
scrubbing process. The calculated avoided cost was above 60 $/ton of CO2. It was
emphasized that public financial support is essential for applications of such
technologies in a cement plant since the current market conditions can only provide
limited financial support.

4.2 Method of Economic Analysis

The economic study reported in this chapter describes the assessment of total
capital requirement, operating & maintenance (O&M) and variable costs of the
carbon capture technologies presented in Chapters 2 and 3, in addition to, the various
sensitivities of cost parameters. The economic analysis was based on the approach
given by IEA (2008). Similar to the levelised cost of energy (LCOE) which is
applicable for power plants (DECC, 2012), levelised cost of cement (LCOC) that
refers the ratio of the net present value of total capital requirement, O&M and

120
variable costs of a cement plant to the net present value of cement production over its
operating life has been calculated. The formulation of the LCOC is given by the
following equation:

𝑇𝐶𝑅𝑡 +𝑀𝑡 + 𝑉𝑡 + (𝑇𝐶𝑅𝑡,𝑐𝑐 +𝑀𝑡,𝑐𝑐 +𝑉𝑡,𝑐𝑐 )


∑𝑛
𝑡=1 (1+𝑟)𝑡
𝐿𝐶𝑂𝐶 = 𝐶 (Eq. 4.5)
∑𝑡=1 𝑡 𝑡
𝑛
(1+𝑟)

where TCRt is the total capital requirement, Mt is the O&M cost, Vt is the variable
cost, and Ct denotes the cement production rate. The suffix cc refers to carbon
capture process, and t and r are the operating year and discount rate, respectively. By
knowing the difference between the LCOC estimations for the base cement plant and
a capture case, and the CO2 emissions for each case, the cost of CO2 avoided can be
calculated according to the following equation:

𝐿𝐶𝑂𝐶𝑐𝑐 −𝐿𝐶𝑂𝐶𝑏𝑎𝑠𝑒
𝐶𝑜𝑠𝑡 𝑜𝑓 𝐶𝑂2 𝑎𝑣𝑜𝑖𝑑𝑒𝑑 = (Eq. 4.6)
[𝑡𝑜𝑛 𝐶𝑂2 ⁄𝑡𝑜𝑛 𝑐𝑒𝑚𝑒𝑛𝑡]𝑏𝑎𝑠𝑒 −[𝑡𝑜𝑛 𝐶𝑂2 ⁄𝑡𝑜𝑛 𝑐𝑒𝑚𝑒𝑛𝑡]𝐶𝐶

The main financial assumptions were taken from IEA (2008) and IEA GHG R&D
programme Technical & Financial Assessment Criteria (IEA GHG, 2003) for a
cement plant located in Scotland, UK and are summarized in Table 4-1. In the
economic analysis of the capture cases, the main issue is the estimation of TCR of
each system. The routine and breakdown maintenance is allowed for at 3.5% per year
of the TCR for the plants handling solids and at 1.7% per year for the plants handling
gases and liquids according to the IEA report (2008). The operating labour is
assumed to work in a 5 shift pattern with an annual salary of €40000/yr. An
allowance of 20% of the operating cost is added to cover supervision and an
additional 30% for administration and general overheads. To cover specific services,
e.g. local rates and insurance, 1.7% of the TCR is included. An escalation factor of
1.5% is included for the estimation of O&M cost during the operating life. Although
a more complex model can be used for financing assumptions (ZEP, 2011), the cost
model used in this dissertation is assumed to be sufficient for the comparison of
different carbon capture technologies.

121
IEA (2008) reported the reference cement plant capital cost allocation, and the
calculations and assumptions have been replicated here for the base cement plant.
The reference capital cost data for the capture processes reported in Table 4-2 were
obtained from the relevant references (DOE, 2003; Fisher et al., 2004; IEA, 2008).
The similarity of the Ca-looping process with an existing CFB oxy-fired power plant
was taken into consideration. The TCR of the reference oxy-fired power plant is
comprised of the costs of a CFB oxy-fired boiler, an ASU, a steam cycle and a gas
processing system including CO2 compression. The total cost was divided under four
titles, as presented in Table 4-2 with their capacities. A further 10% of the TCR of a
complete oxy-fired power plant was added to cover carbonator cost as suggested by
Abanades et al. (2007) and Rodriguez et al. (2012). Since the flow rate of raw meal
into the base cement plant decreases with increasing purge flow rate in the Ca-
looping and Ca-Cu looping processes, the costs of the preheater and pre-calciner
existing in the base cement plant have been discounted with respect to the flow rate
of the raw meal. The preheater and pre-calciner costs were given as 13.6 M€ and 1.1
M€ at 189 ton raw meal/h, respectively (IEA, 2008).

The main components of the TCR in the indirect calcination process are a CFB
combustor and a gas processing system. Because, the first is similar to the boiler of
an air-fired CFB power plant, the TCR estimations given in DOE (2003) for this
plant were divided into two categories: boiler island and steam cycle. The cost of the
CFB combustor in the indirect calcination process was then calculated based on the
total thermal energy requirement in this reactor.

Table 4-1 The main financial assumptions (IEA, 2008).

Value Unit
‘S’ curve of expenditure
1st Year 20 %
2nd Year 45 %
3rd Year 35 %
Design life 25 Years
Load (or capacity) factor
1st Year 60 %
2nd Year 90 %
Escalation factor 1.5 %
Discount factor 10 %
Conversion factor ($ to €) 0.77 -

122
Table 4-2 Reference capital cost data.

Capacity Cost Reference


Reference cement plant 114.9 ton clinker/h 263 M€ (IEA, 2008)
CFB oxy-fired combustor (calciner) 1877 GJth/h* 94.5 M$ (DOE, 2003)
ASU 37.5 MWe 76.8 M$ (DOE, 2003)
Steam cycle 209.9 MWe 85 M$ (DOE, 2003)
Gas processing system 26.9 MWe 72.2 M$ (DOE, 2003)
CFB air reactor/fuel reactor 1845 GJth/h* 122 M$ (DOE, 2003)
CFB air-fired combustor 1884 GJth/h* 164 M$ (DOE, 2003)
Absorber (MEA) 1110 m3 8.3 M$ (Fisher et al., 2004)
Stripper (MEA) 238 m3 1.9 M$ (Fisher et al., 2004)
Reboiler (MEA) 125 MWth 2.4 M$ (Fisher et al., 2004)
CHP 746 GJth/h* 146.4 M€ (IEA, 2008)
FGD 143 m3/s at 50°C 49.8 M€ (IEA, 2008)
SCR 89 m3/s at 320°C 10.2 M€ (IEA, 2008)
ESP 76.8 m3/s at 110°C 4.6 M€ (IEA, 2008)
* LHV Basis

The reference TCR of the CLC system was taken from the similar process
configurations detailed in the report by the DOE (2003). Despite this system’s
similarities to the Ca-looping process, it does not include an ASU. For the amine
process, the cost structure was designed to include the absorber, desorber,
compressors, pumps, heat exchangers and gas processing system. The reference heat
transfer coefficients required sizing the heat exchangers and heat exchanger types in
the amine process were taken from the similar type of work, in addition to the costs
of the absorber, desorber and reboiler at known capacity (Fisher et al., 2004). The
costs associated with the heat exchangers, pumps and compressors were calculated
using the CAPCOST software (Turton et al., 2009). The absorber and stripper were
sized by the Tray Sizing Tool in UniSim based on packed bed column using Flexipac
250Y and IMTP #40, respectively. Since it is not possible to size some auxiliaries in
the amine process such as reflux accumulator, filtration and reclaimer by the process
simulator, the costs of these units were assumed to be 8% of the TCR of the amine
process according to Fisher et al. (2004). The cost of a CHP plant including an SCR
unit was estimated based on the thermal energy demand of this plant. Even though
the costs of the FGD and SCR units were estimated based on volumetric gas flow
rate, it is worth to mention that more accurate predictions can only be achieved if
detailed cost data for such systems are provided.

123
For the membrane process, the contributions in cost of the membrane module,
membrane frame and material replacement can be estimated based on the following
equations:

𝑇𝐶𝑅𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒−𝑚𝑜𝑑𝑢𝑙𝑒 = 𝐴𝑚 𝑐𝑚 (Eq. 4.7)

𝑇𝐶𝑅𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒−𝑓𝑟𝑎𝑚𝑒 = (𝐴𝑚 ⁄2000)0.7 𝑐𝑚𝑓 (Eq. 4.8)

𝑇𝐶𝑅𝑚𝑒𝑚𝑏𝑟𝑎𝑛𝑒−𝑟𝑒𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 = (𝐴𝑚 𝜗)𝑐𝑟𝑚 (Eq. 4.9)

In all three of these equations, the following apply: Am refers to the required
membrane area; cm, cmf and crm are set to 50 $/m2, 0.238 M$ and 10 $/m2,
respectively (IECM, 2012); and the annual material replacement rate (𝜗) is fixed at
20%. In addition, a loading of 40% for contingencies and remaining fees needs to be
included (IECM, 2012). The material replacement cost has been included as an
O&M cost. The costs of compressors, vacuum pumps and expanders in this system
were also calculated by using the CAPCOST software.

According to the six-tenths rule, the approximate capital requirement of a plant or


a unit can be estimated if the cost and size of similar items are known. Moreover,
historical cost predictions can be adjusted to current prices due to inflation and
deflation by using the Marshall&Swift (M&S) index. The following equation
demonstrates the use of the six-tenths rule and the M&S index:

𝐶 0.6 𝑀&𝑆𝐵
𝑇𝐶𝑅𝐵 = 𝑇𝐶𝑅𝐴 × � 𝐵 � ×� � (Eq. 4.10)
𝐶𝐴 𝑀&𝑆𝐴

where TCRA is the known cost of an item with size CA; TCRB is the estimated cost of
the equipment with size CB; and the term M&S refers to the index value on a base
year. Therefore, Eq. 4.10 has been applied on the available capital cost data in order
to estimate current prices. An exception is given for the membranes for which scale
factor is assumed to be 1 since they are modular. The fuel and raw material costs
were taken from IEA (2008) and are summarized in Table 4-3. These values have
been updated using 1.5% escalation factor per year to estimate current prices (ZEP,
2011).

124
Table 4-3 Reference variable cost data. * The cost of SCR catalyst has been adjusted based on the
actual SCR capacity in each case.

Cost Unit Reference


Raw material + Process water 6.1 €/ton cement (IEA, 2008)
Limestone for FGD 3.0 €/ton (IEA, 2008)
Ammonia 2000 €/ton (IEA, 2008)
MEA 1100 €/ton (IEA, 2008)
SCR catalyst 1.19* M€/year (IEA, 2008)
Pet coke 80.0 €/ton (IEA, 2008)
Coal 65.0 €/ton (IEA, 2008)
Electricity 0.05 €/kWh (IEA, 2008)
Copper 6.8 $/kg (LME, 2013)
Alumina 0.8 $/kg (INDMIN, 2013)
Natural Gas 4.9 $/1000 ft3 (EIA, 2013)

The electricity cost was set to 0.05 €/kWh. As a base approach, the same
electricity cost was utilized as revenue for surplus power generation (IEA, 2008;
Rodriguez et al., 2011b). An average annual escalation factor of 1.5% has also been
considered in all variable cost calculations for the subsequent years. To estimate the
cost of CuO/Al2O3 sorbent in the Ca-Cu looping process, the following equation
proposed by Abad et al (2007) has been employed.

𝐶𝑂𝐶 = 𝑥𝐶𝑢𝑂 𝐶𝐶𝑢𝑂 + 𝑥𝐴𝑙2𝑂3 𝐶𝐴𝑙2𝑂3 + 𝐶𝑚 (Eq. 4.11)

where x and C refer to component mass fractions and cost, respectively for oxygen
carrier (OC) and its constituents; CuO and Al2O3. The approximate manufacture cost
of oxygen carrier, Cm, was given as 1 $/kg OC in Abad et al (2007).

4.3 Results and Discussion

4.3.1 Base Cement Plant and Ca-looping Process

The calculated LCOC estimates presented in Figure 4-1 are for the base cement
plant, Ca-looping options at three different F0/FCO2 ratios providing >90% CO2
avoidance (see Figure 2-8), and the oxy-calciner only case from Chapter 2. A sample
economic analysis spreadsheet for the Ca-looping process at 1.65 F0/FCO2 is attached
in Appendix C.

125
160

120

LCOC [€/ton cement]


ETS
Variable
O&M
80
TCR

40

0
Base Cement 0.2 1.65 3 Oxy-calciner
Plant Only
Figure 4-1 LCOC estimates for the base cement plant, three Ca-looping options at different F0/FCO2
(see Figure 2-8) and oxy-calciner only case. The error bar reflects the increase in variable cost if there
is no surplus power revenue.

Table 4-4 The composition of the TCRs for the Ca-looping and ‘oxy-calciner only’ processes
presented in Figure 4-1.

0.2 1.65 3.0 Oxy-calciner


Unit
(×106 €) (×106 €) (×106 €) (×106 €)
Calciner 51.2 45.9 46.1 44.0
Steam Cycle 45.9 33.0 31.2 26.7
ASU 42.0 37.7 37.8 36.1
Gas processing system 57.6 52.8 52.5 48.9
Carbonator 19.7 16.9 16.8 -
Cyclone 1.9 1.2 1.0 -
Total 218.3 187.5 185.4 155.7
Upon further consideration, a significant difference between the TCR of the base
cement plant and that of the capture options can be observed. The increment reaches
a maximum at 0.2 F0/FCO2 due to the excessive thermal energy requirement in this
system. The composition of TCR estimations for the capture plants is given in Table
4-4. In parallel with a reduction in the thermal energy requirement, the TCR
estimates decrease by around 15% for 1.65 F0/FCO2 and 3.0 F0/FCO2 cases.

Although the cost of the calciner is the lowest at 1.65 F0/FCO2 among the Ca-
looping options owing to the minimum thermal energy requirement in this system as
presented in Figure 2-6(a), the lowest TCR estimate belongs to the oxy-calciner only
case mainly because of the absence of the costs of the carbonator and additional
cyclone. The O&M estimates are directly related to the TCR estimates so a more
pronounced effect is observed at 0.2 F0/FCO2. The main components of the variable

126
cost in the base cement plant and Ca-looping processes are the costs of raw material,
fuel and electricity. The base plant should pay for the electricity as it is imported,
while in the other cases, either all (0.2 F0/FCO2) or partial (1.65 and 3.5 F0/FCO2, and
the oxy-calciner only case) requirement can be met by on-site power generation. As
indicated previously, the surplus power generation at 0.2 F0/FCO2 has been included
as revenue in the cost model. However, the surplus power revenue will also vary on a
location basis: if for example the infrastructure for the electricity grid is not able to
take large amounts of electricity from the cement plant, the cement plant will not be
able to sell the surplus electricity. Therefore, an error bar is also included in Figure 4-
1 to reflect the case where surplus power generation cannot be considered as
revenue, increasing the variable cost further.

The variable cost of the Ca-looping process at 0.2 F0/FCO2 which is the case with
the greatest power generation is slightly lower, despite having the highest cost
associated with fuel consumption. The trade-off between the fuel and electricity costs
keeps the variable costs of the 1.65 F0/FCO2 and 3.0 F0/FCO2 cases at the same level
while further increase is observed in the oxy-calciner only case. To reveal the
potential advantages of cement plants with CO2 capture, the additional benefits from
the Emission Trading System (ETS) is included in the economic analysis. According
to the ETS, industries emitting CO2 pay a variable amount for the emission credits,
depending on market conditions. As a reference value of this emission credit, 14
€/ton CO2 has been adopted as this was the value used in the IEAGHG report (2009).
The contribution of the ETS price is the highest for the base cement plant since all
CO2 generated in this system is directly emitted whereas it is significantly less for the
Ca-looping processes where more than 90% CO2 avoidance can be achieved. At 3.0
F0/FCO2, the minimum ETS price is utilized since the highest CO2 avoidance rate has
been achieved in this case (see Figure 2-8). Even at the different F0/FCO2 ratios, the
LCOC estimates for the Ca-looping process remains constant at approximately 127
€/ton cement (140 €/ton excluding surplus power revenue at 0.2 F0/FCO2), while that
of the base cement and the oxy-calciner only plants was calculated to be 97 €/ton
cement and 124 €/ton cement, respectively.

127
65
>90%
>90% (no surplus power revenue)
90%
90% (no surplus power revenue)
Oxy-calciner only

[€/ton CO2 avoided]


Cost 55

45

35
0 1 2 3 4 5 6
F0/FCO2

Figure 4-2 The cost estimates for the Ca-looping and ‘oxy-calciner only’ processes. For the Ca-
looping process, the CO2 avoidance rate is either fixed at ‘90%’ or ‘>90%’, reflecting the results given
in Chapter 2.

The use of LCOC results in Eq. 4.6 enables the calculation of the cost of CO2
avoided. The results shown in Figure 4-2 are for the Ca-looping process at two
different CO2 avoidance rates; (i) 90% (see Figure 2-9) and (ii) >90% (see Figure 2-
8), and for the oxy-calciner only case. Given the fact that the revenue of surplus
power generation depends on demand in the grid, additional results have been
included for the cases at <1.5 F0/FCO2, assuming no surplus power revenue. The cost
of CO2 avoided is the highest at the lowest F0/FCO2 owing to the greater thermal
energy and oxygen requirements and reduces if the avoidance rate is more than 90%.
The main difference between the ‘90%’ and ‘>90%’ cases is the amount of sorbent
circulated between the reactors which affects the heat requirement in the calciner,
whereas the volumetric flow rate of the feed flue gas from the cement plant as well
as the purge flow rate is almost similar. In addition, the crucial term in the cost of
CO2 avoided calculations (Eq. 4.6) is the CO2 intensity which decreases as the CO2
avoidance rate increases.

128
The cost of CO2 avoided estimates at 90% CO2 avoidance show a similar trend to
the heat requirement in the calciner as presented in Figure 2-6(a), while the others at
>90% CO2 avoidance stay almost constant after 0.4 F0/FCO2 with a value of 41 €/ton
CO2 avoided. The cost associated with the oxy-calciner only case is around 1 €/ton
CO2 avoided lower than that of the Ca-looping plant (minimum 2 €/ton CO2 avoided
at 90% case). Without any revenue for surplus power generation, a severe increment
in the cost estimates is observed especially at 0.2 F0/FCO2 where the cost reaches 62
€/ton CO2 avoided.

4.3.2 Ca-Cu Looping Process

Even though the Ca-Cu looping process provides an improvement in the energetic
efficiency, the major drawback on the process economics results from the cost of the
oxygen carrier that is purged from the system. The LCOC estimates for the Ca-Cu
looping process and the Ca-looping process using methane as fuel in the calciner are
presented in Figure 4-3 with error bars reflecting the conditions with no surplus
power revenue. The breakdown of TCR calculations for the capture plants is
illustrated in Table 4-5. Here, cases A and B refer to the Ca-looping process using
natural gas as a fuel at 0.02 F0/FCO2 and 0.15 F0/FCO2, respectively while the same
purge ratios refer to cases C and D for the Ca-Cu looping process. The CO2
avoidance rate in both processes has been fixed at 90%. Predictably, the TCR
estimates for the cases A and B are lower compared to equivalent configurations at
low F0/FCO2 burning pet-coke (see Figure 4.1). This occurs because of the reduction
in thermal energy requirements of the calciner in the cases A and B as a result of the
absence of negative SO2 and ash effects on the sorbent performance. The elevated
thermal energy requirement increases the reactor and steam cycle costs and leads to
the highest TCR found in case C, even though an ASU does not exist in this system.
It is convenient to assume that the cost of the carbonator in case D should be similar
to that in case B as both systems have similar thermal energy requirements in the
capture plants, CO2 and sorbent flow rates as well as heat exchange area
requirements in the carbonators according to Table 3-4. Thus, to estimate the cost of
the carbonator in cases C and D, a further 11.2% of the TCR of the CLC plant has
been added to be consistent with that of the Ca-looping processes.

129
240

200

LCOC [€/ton cement]


160 ETS
Variable
O&M
120
TCR

80

40

0
A B C D

Figure 4-3 The LCOC estimates for the Ca-looping process using methane as a fuel (cases A and B)
and the Ca-Cu looping process (cases C and D). Cases A and C refer to the results at 0.02 F0/FCO2
while cases B and D refer to those at 0.15 F0/FCO2. The error bars reflect the increase in the variable
cost if there is no surplus power revenue.

Table 4-5 The composition of the TCRs for the Ca-looping process using methane (cases A and B)
and the Ca-Cu looping process (cases C and D) presented in Figure 4-3.

A B C D
Unit
(×106 €) (×106 €) (×106 €) (×106 €)
Calciner 44.6 42.8 - -
Steam Cycle 42.3 38.0 58.3 40.4
ASU 36.7 35.2 - -
Gas processing system 49.9 48.3 53.0 47.2
Air Reactor/Fuel Reactor - - 80.4 59.0
Carbonator 17.4 16.4 21.5 16.4
Cyclone 2.1 2.0 2.1 2.0
Total 193.0 182.7 213.0 163.3

Notably, when the 0.02 F0/FCO2 cases are compared, it can be realized that
regardless of the excessive cost of oxygen carrier in the Ca-Cu looping process, the
estimated variable cost is close to that of the Ca-looping case. It can be associated to
very small flow of CuO in the purge. However, without any revenue of surplus
power generation, the LCOC estimates increase in all cases, particularly in Case C.
The cost of sorbent is very prominent at in case D, and this option is not competitive
with the other cases.

130
The LCOC estimates are in the range of 130 €/ton cement to 138 €/ton cement for
most of the cases but reaches a maximum of 201 €/ton cement in case D including
the ETS price. Without any surplus power revenue, a severe jump in the cost
estimate of case C can be marked, indicating the importance of the export of surplus
power for this case. The cost of CO2 avoided estimates for the Ca-looping process
using methane and the Ca-Cu looping process are presented in Figure 4-4. The
results are divided into four different options: the first option (Option 1) directly
represents the conditions applied in Figure 4-3; the second option (Option 2) shows
the electricity selling price has increased to 0.1 €/kWh which is an advantage for the
processes generating great amount of electricity; the third option (Option 3) is
opposite to Option 2 where there is no revenue of surplus power generation; and the
fourth option represents the best scenario applicable for the Ca-Cu looping process
where entire purged CuO/Al2O3 sorbent is separated and reused in the system.

The cost estimates for cases A and B according to Option 1, are around 50 €/ton
CO2 avoided; these values are slightly higher than the previous ‘90%’ case using pet-
coke as presented in Figure 4-2 because of the higher cost of natural gas than pet-
coke. At 0.02 F0/FCO2, the Ca-Cu looping process is more competitive compared to
the Ca-looping options, but the sorbent cost becomes challenging at 0.15 F0/FCO2
where the cost increases to 157 €/ton CO2 avoided. Additionally, the electricity
selling price impacts the total variable cost estimates, and this value is kept at 0.05
€/kWh initially but changed to 0.1 €/kWh in Option 2. For the processes importing
the electricity, as in the base cement plant, an increase in the electricity cost triggers
the variable costs. However, for the capture processes where power generation is
possible, the increase causes a reduction in the variable cost. The profit is more
obvious for the cases at 0.02 F0/FCO2 since the power generation capacities are
higher. Because the power generation capacity is the greatest in case C, one should
expect severe variation in the cost of this system compared to the others. For this
case, the cost of CO2 avoided takes a negative value. Conversely, for an opposite
scenario presented in Option 3, the cost increases to 114 €/ton CO2 avoided in case
C.

131
180
150

[€/ton CO2 avoided]


Option 1
120 Option 2
Cost
90 Option 3

60 Option 4

30
0
A B C D
-30

Figure 4-4 The cost of CO2 avoided estimates for the Ca-looping process using methane (A and B)
and the Ca-Cu looping process (C and D).

Although it was not considered in the process simulations presented in Chapter 3


and the CuO content in the clinker is limited to 1 wt%, the possible recovery of the
CuO/Al2O3 sorbent leaving in the purge stream will reduce the variable cost
significantly. In Option 4, the reduction in the cost of CO2 avoided is more
prominent in case D. It decreases to 38 €/ton CO2 avoided exclusive of cost of solid
separation. The results suggest that if there is an economical way to separate
CuO/Al2O3 from the solid mixture, the Ca-Cu looping process will be more
competitive with the Ca-looping process in terms of process economics.
Furthermore, this will also allow the operation of the system with high F0/FCO2,
reducing the cost of CO2 avoided further.

4.3.3 Indirect Calcination, Hybrid and Standalone Amine Processes

The TCR estimates for the standalone indirect calcination, hybrid and standalone
amine processes are presented in Table 4-6. The TCR of the standalone indirect
calcination process is comprised of the costs of a CFB combustor, a gas processing
system and a regenerative heat exchanger to heat up the excess air stream by heat
exchange with the CO2-rich gas stream. The total TCR of the standalone amine
process is significantly higher than that of the base cement plant, whereas the hybrid
configuration provides around a 30% improvement in the TCR even though both
processes provide 90% CO2 avoidance. It should be noted that the TCRs of the
hybrid and standalone amine processes include also the costs of the FGD and SCR

132
units, causing reduction of SOx and NOx in the final gas stream, while these units do
not exist in the standalone indirect calcination process.

The LCOC estimates for these systems are shown in Figure 4-5. Similarly, an
error bar indicates the increase in the variable cost if there is no revenue for surplus
power generation. The cement cost calculated for the standalone indirect calcination
process is only 12 €/ton cement higher compared to the base cement plant. Although
the TCR and O&M costs are greater for the indirect calcination process, the total
difference reduces with the inclusion of the ETS price.

133
Table 4-6 The composition of the TCRs for the standalone indirect calcination, hybrid configuration and standalone amine process.

Standalone indirect Hybrid Standalone amine


calcination configuration process
(×106 €) (×106 €) (×106 €)
Amine Process
Flue gas blower - 2.3 4.5
Absorber - 7.4 14.1
Rich amine pump - 0.04 0.09
Stripper - 1.3 3.2
Reflux condenser - 0.8 2.0
Reboiler - 1.3 2.9
Rich/lean amine heat exchanger - 5.9 10.0
Lean amine pump - 0.005 0.04
Lean amine cooler - 0.3 3.7
Inlet gas cooler - 1.6 2.9
Other auxiliaries - 2.3 3.8
CFB combustor 53.2 - -
CHP plant - 120.8 170.8
SCR unit - 4.3 10.8
FGD unit - 38.5 55.4
Gas processing system 30.2 46.2 53.8
Regenerative heat exchanger 0.5 0.5 -
Total 83.9 233.0 338.0
134
200

160

LCOC [€/ton cement]


ETS
120 Variable
O&M
TCR
80

40

0
Standalone indirect Hybrid configuration Standalone Amine
calcination Process

Figure 4-5 The LCOC estimates for the standalone indirect calcination, hybrid configuration and
standalone amine process. The error bar reflects the increase in variable cost if there is no surplus
power revenue.

The LCOC increases to 146 €/ton cement for the hybrid process and 168 €/ton
cement (172 €/ton cement without the surplus power revenue) for the standalone
amine process. Even though there is no power generation facility in the standalone
indirect calcination process, its variable cost estimation is very low due to the
absence of excessive costs of fuel and solvent.

The cost of CO2 avoided estimates are shown in Figure 4-6. For the standalone
indirect calcination process, the estimated cost is around 29 €/ton CO2 avoided,
which is the lowest among the different options compared in this dissertation despite
the fact that the CO2 avoidance rate provided by this process is only 56%. By adding
an amine process, 90% CO2 avoidance can be attained causing the cost to increase to
73 €/ton CO2 avoided, which is up to 80% higher than that of the Ca-looping process.
However, compared to the standalone amine process, it provides more than a 30%
improvement in the cost.

135
120

100

[€/ton CO2 avoided]


80

Cost
60

40

20

0
Standalone indirect Hybrid configuration Standalone amine process
Calcination

Figure 4-6 The cost of CO2 avoided estimates for the indirect calcination, hybrid configuration and
standalone amine process. The error bar reflects the case where there is no revenue for surplus power
generation.

4.3.4 Membrane Process

For the membrane process, the impact of the predefined range of S/R ratios on the
LCOC estimates has initially been examined, and the results are presented in Figure
4-7. The cost results are given for two cases: in Option 1, the feed stream is the gas
stream leaving the preheater while it is the gas stream after the bag filters (end-of-
pipe) in Option 2. It can be concluded that even though distinctive values of the
membrane areas and energy consumptions have been calculated at different S/R
ratios, as already shown in Figure 3-18, the effect is almost negligible on LCOC
estimates. This occurs because of the trade-off between the costs of membrane
module, vacuum pumps and compressors. The LCOC estimates are close at 0.1 S/R
and 0.2 S/R but around 1 €/ton cement higher at 0.05 S/R. Despite the membrane
area is 48% higher in Option 2; the LCOC predictions based on two feed locations
are not significantly different from each other.

The breakdown of TCR estimates for the membrane process based on Options 1
and 2 at 0.1 S/R is given in Table 4-7. The total cost of the heat exchangers includes
the one needed in Option 1 to heat the CO2-depleted gas stream. Therefore, only the
total heat exchanger cost is greater in Option 1 while all remaining are higher in
Option 2 owing to larger membrane area requirement and greater volumetric flow
rate of the feed gas.

136
140

120

LCOC [€/ton cement]


100
ETS
Variable
80
O&M
60
TCR

40

20

0
Option 1 Option 1 Option 1 Option 1 Option 2
(0.05) (0.075) (0.1) (0.2) (0.1)

Figure 4-7 The LCOC estimates for Options 1 and 2 at different S/R ratios. The S/R ratios are shown
in the parentheses.

Table 4-7 The composition of the TCRs for the membrane process at two different feed gas locations.

Option 1 Option 2
(×106 €) (×106 €)
Total membrane (module + frame) 22.7 32.3
Heat exchangers 1.2 0.8
Compressors, vacuum pumps and expanders 6.9 8.2
SCR 10.8 10.8
FGD 27.7 34.5
Gas processing system 39.1 39.1
ESP 3.9 -
Total 112.3 125.7

The cost estimates for the membrane process are compared in Figure 4-8 for
Options 1 and 2. With this technology, the cost can be reduced to approximately 42
€/ton CO2 avoided if the feed gas location is selected as the exit of the preheater. If a
complete retrofit option is preferred, the cost will increase to 47 €/ton CO2 avoided.
The cost estimate for the membrane process taking the feed stream from the exit of
the preheater is up to 1 €/ton CO2 avoided higher than that of the Ca-looping process
as presented in Figure 4-2. Even though the TCR and O&M cost are higher for the
latter, the variable cost requirement in this system is lower because of the on-site
power generation. However, the footprint of the membrane process would be smaller
so it may be more attractive in the case of retrofits to existing plants with limited
availability of land.

137
50

40

[€/ton CO2 avoided] 30


Cost

20

10

0
Option 1 Option 2

Figure 4-8 The cost of CO2 avoided estimates for the membrane process at two different feed gas
locations.

4.3.5 Sensitivity Analysis


In the final stage of this chapter, sensitivities of TCR, O&M cost, fuel and power
costs, discount rate, scale factor and ETS cost based on a variation of ±25% have
been examined. In addition, three additional scenarios, where (i) CO2 saving by on-
site power generation, (ii) inclusion of FGD and SCR units in all process schemes,
and (iii) replacement of the ASU in the Ca-looping process with oxygen purchasing
have been taken into the consideration. The sensitivity analysis has been developed
for seven different carbon capture options: (i) Ca-looping process at 1.2 F0/FCO2 (the
case where the CO2 avoidance rate is ≥90% and the entire electricity requirement can
be met by on-site power generation with an excess of less than 2 MWe), (ii) Ca-Cu
looping process at 0.02 F0/FCO2, (iii) standalone indirect calcination process, (iv)
hybrid configuration, (v) standalone amine process, (vi) membrane process (at 0.1
S/R, Option 1), and (vii) oxy- calciner only case. First, the results for the sensitivities
of TCR and O&M cost are shown in Figure 4-9.

138
(a)
140

120

[€/ton CO2 avoided]


100
Cost
80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

(b)
140

120
[€/ton CO2 avoided]

100
Cost

80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

Figure 4-9 Sensitivities of (a) TCR and (b) O&M.

While the variation in TCR is greater for the standalone amine process as the cost
estimate for this system is the highest, it changes in a small range for the standalone
indirect calcination and membrane processes. In Figure 4-9(b), the sensitivities of
O&M cost indicate that similar to the variation in the TCR, the fluctuation is severe
for the Ca-Cu looping process, hybrid configuration and standalone amine process.

139
(a)
140

120

[€/ton CO2 avoided]


100
Cost
80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

(b)
140

120
[€/ton CO2 avoided]

100
Cost

80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

Figure 4-10 Sensitivities of (a) fuel cost and (b) power cost.

The sensitivities of fuel and power costs are presented in Figure 4-10. In Figure 4-
10(a), there is not a significant change in the costs of the indirect calcination process
and the membrane process, since the total thermal energy requirement in these
systems are similar to that of the base cement plant. A change in the fuel cost affects
the variable cost in both the base cement plant and these capture processes, therefore,
this effect disappears for the cost of CO2 avoided estimates because the increase in
the cost is similar. A more pronounced effect is observed for the Ca-Cu looping
process and the standalone amine process due to the excessive thermal energy
requirements in these systems.

140
The impact of a change in power cost, shown in Figure 4-10(b), is different for the
capture cases. There is power generation in the majority of the configurations except
the indirect calcination and membrane processes. For these exceptions, an increment
in the electricity cost causes the cost of CO2 avoided to increase, while the converse
occurs for the remaining processes. Although the CHP plant located in the hybrid
configuration provides electricity to be used in the system, the final electricity cost is
still higher than that of the base cement plant. Therefore, it is preferable for this
option to pay less for the electricity. Conversely, in the oxy-calciner only case, the
need for electricity is lower than that in the base cement plant, so the cost of CO2
avoided is lower at high electricity cost.

The sensitivities of discount rate and the scale factor that is assumed to be 0.6
initially are presented in Figure 4-11. In parallel to an excessive TCR estimate for the
standalone amine process, the cost of CO2 avoided calculated for this system ranges
between 95 €/ton CO2 avoided to 119 €/ton CO2 avoided based on ±25% variation of
the discount rate. For the remaining cases up to 30% variation is observed depending
on the severity of TCR estimates. The sensitivity on scale factor, presented in Figure
4-11(b), is only applicable to the processes for which the cost estimations are based
on Eq. 4.10. Even though the costs of the absorber, desorber, reboiler and gas
processing unit existing in the hybrid and standalone amine processes have been
calculated by using the same equation, the costs of the compressors, expanders and
heat exchangers in these systems have been taken from the CAPCOST software. For
this reason, the results shown in Figure 4-11(b) are only for the Ca-looping, Ca-Cu
looping process, indirect calcination and the oxy-calciner only processes. For the
selected systems, an increment in the scale factor reduces the cost of CO2 avoided
since the reference cost data for these systems is at higher capacity, leading the cost
to be reduced to 22 €/ton CO2 avoided for the indirect calcination process and to 33
€/ton CO2 avoided for the Ca-looping process.

141
(a)
140

120

[€/ton CO2 avoided]


100
Cost
80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

(b)
140
120
[€/ton CO2 avoided]

100
Cost

80
60
40
20
0
Ca-looping Ca-Cu looping Indirect Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Only

Figure 4-11 Sensitivities of (a) discount rate and (b) scale factor.

Finally, the sensitivity of the cost of CO2 avoided estimates against the ETS cost
is given in Figure 4-12. An increase in the ETS cost is necessary to make the capture
processes competitive with the base cement plant as it allows a reduction in the cost
of CO2 avoided. An additional 25% increase in the ETS cost reduces the cost of the
Ca-looping process to 37 €/ton CO2 avoided while that of the indirect calcination
process reduces to 25 €/ton CO2 avoided. If the ETS cost is set to 50 €/ton CO2, the
cost model gives almost zero cost of CO2 avoided estimate for the Ca-looping case,
which will facilitate the use of Ca-looping process in the cement industry.

142
140

120

[€/ton CO2 avoided]


100

Cost
80

60

40

20

0
Ca-looping Ca-Cu looping Indirect Hybrid Standalone Membrane Oxy-calciner
(1.2 F0/FCO2) (0.02 F0/FCO2) calcination Configuration Amine Only

Figure 4-12 Sensitivity of ETS cost.

4.3.5.1 The Effect of Grid Emission Factor

The emission factor for average grid electrical consumption refers to the CO2
emitted during electricity generation. An average value of 520 kg/MWh is invoked as
a reference value in the course of this analysis; this value has been derived from the
IEA (2008), which appropriated the emission factors supplied by DEFRA, as well as
the energy statistics that were reported in the Digest of UK Energy Statistics
(DUKES).

In particular, the emissions from coal, oil, gas, nuclear, renewable and other types
of energy and energy generating mechanisms, such as wind turbines, were divided by
the total electricity supplied to the grid. The use of the grid emission factor in the
calculations will highlight the CO2 emissions in the base cement plant; it will also
emphasise options where electricity is imported. To simplify the use of the grid
emission factor, negative emissions that are generated as a result of surplus power
generation are not taken into account. For example, in cases such as the Ca-looping
process, the Ca-Cu looping process and the standalone amine process, wherein there
is surplus power generation, the CO2 emissions that relate to the use of electricity is
set to zero instead of considering a negative value.

143
Table 4-8 The cost of CO2 avoided estimates using the grid emission factor.

Cost
(€/ton CO2 avoided)
Ca-looping (1.2 F0/FCO2) 38.0
Ca-Cu looping (0.02 F0/FCO2) 56.1
Indirect calcination 31.5
Hybrid configuration 75.6
Standalone amine process 98.1
Membrane process (0.1 S/R) 47.1
Oxy-calciner only 38.7

Table 4-8 presents the cost of CO2 avoided estimates, which have been modified
to include the grid emission factor. The table clearly illustrates that the cost decreases
to 38 €/ton CO2 avoided for the Ca-looping option. Similarly, the cost estimates are
reduced for the Ca-Cu looping, standalone amine and oxy-calciner only processes.
Nevertheless, the cost estimate increases to around 47 €/ton CO2 avoided for the
membrane processes as the electricity is imported in this case.

4.3.5.2 Inclusion of FGD and SCR Units in All Configurations

In the cement industry, SO2 is formed due to the combustion of fuel in the pre-
calciner and kiln, and oxidation of sulfur in raw meal in the preheater stage.
Although the SO2 formed by fuel combustion is mostly absorbed by feed materials in
the process, part of the SO2 generated by the raw meal oxidation is emitted. Based on
the given raw meal composition in Table 2-3 where there is only 0.3 wt% sulfur in
the raw meal and following the design basis in IEA (2008), FGD and SCR units have
not been included in the base cement plant described in Chapter 2. However, these
units should be included when amine and membrane separation processes are
retrofitted to prevent the negative effects of SOx and NOx in the feed gas stream.
Various solutions exist to keep these impurities in cement flue gases below the
environmental limits even for S-rich raw meal as explained in details in IEA (2008).
However, to give more flexibility in terms of fuel selection, and considering that the
environmental regulations will become more stringent in the future, the cost results
have been updated to include FGD and SCR units in all capture processes including
the base cement plant.

144
Table 4-9 Variation in the cost of CO2 avoided estimates when the capital costs of FGD and SCR
units are included in all configurations.

Cost
(€/ton CO2 avoided)
Ca-looping (1.2 F0/FCO2) 38.0 (-8%)
Ca-Cu looping (0.02 F0/FCO2) 60.0 (-2%)
Indirect calcination 35.8 (+22%)
Hybrid configuration 55.5 (-24%)
Standalone amine process 89.2 (-16%)
Membrane process (0.1 S/R) 24.4 (-41%)
Oxy-calciner only 33.0 (-17%)

The updated cost estimates including their variation against the previously
reported values in Figures 4-2, 4-4, 4-6 and 4-8 are presented in Table 4-9. Since the
TCR of the base cement plant also increases due to the additional costs of FGD and
SCR units, the cost of CO2 avoided estimates reduces in majority of the cases except
the indirect calcination process. In this process, the volumetric flow rate of the flue
gas stream fed to the FGD and SCR units is greater as a result of using part of excess
air for the combustion of fuel and raw meal preheating. Part of the gas stream is
diverted to the carbonator in the Ca-looping and Ca-Cu looping processes where SO2
can be removed by reaction with CaO, and also NOx emissions can be minimized in
the CFB combustion systems (DOE, 2003). Thus, the cost estimates of these capture
processes including that of the oxy-calciner only case reduce. In the base cost
scenario, TCRs of the hybrid configuration, standalone amine and membrane
processes already include the cost of FGD and SCR units. For this reason, the cost of
CO2 avoided reduces in these cases, providing the minimum cost of 24.4 €/ton CO2
avoided for the membrane process.

Following the same argument, a different scenario can be applicable to membrane


and amine processes, where the latest developments in this field provide opportunity
to manufacture membranes or solvents having resistance against impurities (SOx,
NOx, etc.) in the feed gas stream. Unless the environmental regulations are so strict,
there will not be any requirement for the FGD and SCR units in this scenario.
Therefore, this will also reduce the cost associated with these systems and possibly
favour the membrane process over the other options for the reduction of CO2
emissions from a cement plant.

145
50

[€/ton CO2 avoided]


Cost

40

Oxygen Purchase
On-site ASU
30
0 10 20 30
Oxygen Sale Price [€/ton O2]

Figure 4-13 Variation in the cost of CO2 avoided estimate of the Ca-looping option when the ASU
existing in the system is replaced by oxygen purchase.

4.3.5.3 Replacement of the ASU in the Ca-looping Process with


.Oxygen Purchase

The oxygen requirement in the Ca-looping process can also be supplied by


purchasing oxygen product instead of having an on-site ASU. To reflect such an
option, the cost associated with the ASU has been subtracted from the TCR estimate
of the Ca-looping process, and cost of oxygen has been added in place as a variable
cost. It was reported that the oxygen production cost can be reduced to 15.5 $/ton O2
compared to DOE/NETL target of 20 $/ton O2 (Alptekin et al., 2008). A range of (10
– 25 €/ton O2) oxygen selling prices have been implemented in the cost model, and
the results are illustrated in Figure 4-13. The cost of CO2 avoided estimate for a
system with the ASU is also included for a comparison. The avoided cost is slightly
lower when the oxygen selling price is set to 10 €/ton O2 compared to having an
ASU in the system. The cost estimates are almost similar when the selling price
increases to 15 €/ton O2 (around 20 $/ton O2), while further increase in the oxygen
price makes the ASU option more economical.

146
4.4 Concluding Remarks

The costs of the CO2 capture technologies evaluated in Chapters 2 and 3 have
been reported. A detailed cost model based on LCOC estimations has been proposed
for the comparison of these processes in terms of cost of CO2 avoided. Even though
the cost data for the capture units are scarce since most of these technologies are still
in a development stage, consideration of their similarity with the existing power
systems is the key assumption in the cost model, which allows a reliable estimate.
The reference cost data have been updated using the six-tenths rule and M&S index
to take the capacity of the unit and historical inflation/deflation rates into account.
The ETS price has also been included in the final cost comparisons to expose the
advantage of carbon capture from cement plants. The cost and process efficiency
estimates for the capture technologies reported in Chapters 2 and 3 are summarized
in Table 4-10 for ease of discussion.

Among various options where the F0/FCO2 ratio has been changed in the Ca-
looping process, there was no significant effect on costs estimates after 0.4 F0/FCO2.
The TCR estimates reduce when the CO2 avoidance rate is fixed at 90% compared to
the case where it is above 90%, but the specific cost per ton of CO2 avoided was
found to be lower for the latter. The cost was estimated to be around 41 €/ton CO2
avoided for the Ca-looping process, and it was around 1 €/ton CO2 avoided lower for
the oxy-calciner only case, but the first can provide a higher CO2 avoidance rate as
indicated in Table 4-10. In some circumstances where the assumption of revenue for
surplus power generation is not applicable, the avoided cost estimates severely
increase below 1.5 F0/FCO2 while the others remain constant. Oxygen purchase with a
cost of ≤15 €/ton O2 was found to be an alternative option to meet the oxygen
demand in the Ca-looping process in place of an on-site ASU.

The Ca-Cu looping process offers a reduction in TCR compared to equivalent Ca-
looping process, which is mostly related to the absence of an ASU. However, the
sorbent cost is the main contributor to the cost estimations. The calculated LCOC for
the Ca-Cu looping process is significant at high purge rates, but gets close to that of
the Ca-looping process if the purge rate is kept low. Despite this, the cost of CO2

147
avoided estimate for the 0.02 F0/FCO2 case is still slightly higher than that of the
equivalent Ca-looping process. Further reduction in the cost of the Ca-Cu looping
process can be achieved by either an increase in the revenue of power generation or
separating CuO/Al2O3 sorbent in the purge for reuse.

The standalone indirect calcination process requires the minimum TCR among
different carbon capture options compared in this dissertation because this system
does not include an additional capture unit but only a gas processing system. The
hybrid configuration, combining the indirect calcination and amine processes, results
in minor gains to the LCOC estimate, compared to the standalone application of the
amine process. Despite this, the final cost of this system is still greater than that of
the Ca-looping process.

It was shown that the impact of S/R ratio on cost estimates is negligible for the
membrane process as a result of a trade-off between membrane area and energy
consumption. The LCOC estimates based on the two feed locations considered were
close; with the cost of CO2 avoided for the membrane process is slightly higher than
the value calculated for the Ca-looping process. It seems to be more advantageous to
apply the Ca-looping if the revenue of power generation is high, whereas in contrary
cases, the membrane process will be the option. The possible future developments of
membrane systems with high resistance against SOx and NOx in flue gases would
eliminate the requirement of FGD and SCR units and further improve the cost of this
process.

The sensitivity analysis on the economic indices reveals that a change in the fuel
cost results in vast fluctuations in the cost estimations especially for the Ca-Cu
looping process and the standalone amine process, for which the energy requirement
is severe. When the power cost deviates by ±25%, the cost of the Ca-Cu looping
process at 0.02 F0/FCO2 varies significantly because of larger surplus power
generation in this system. If the surrounding area has electrical needs and there is a
possibility to sell the surplus electricity for profitable prices, this option would be
favourable.

148
Table 4-10 The summary of cost and process efficiency estimates for the capture processes evaluated in Chapters 2 and 3. The values in the parentheses reflect the case where there is
no surplus power revenue.

Variable Cost
CO2 avoidance Energy consumption TCR O&M Cost inc. ETS Cement Cost Cost of CO2 avoided
rate (%) (GJth/ton CO2 avoided) (€/ton cement) (€/ton cement) (€/ton cement) (€/ton cement) (€/ton CO2 avoided)
Base plant - - 34.8 26.1 36.5 97.4 -
Ca-looping 90-99 2.5 – 3.0 55.6 – 61.8 41.4 - 44.6 21.2 – 29.3 (33.3) 125.7 – 127.5 (139.7) 41.0 – 45.0 (61.7)
Oxy-calciner only 90 2.3 53.2 40.1 30.5 123.8 40.0
Ca-Cu looping 90 1.7 – 2.2 55.8 – 62.3 41.4 – 45.0 29.3 – 102.9(116) 137.6 – 201.3 (214.4) 60.8 – 157.0 (176.6)
Indirect Cal. 56 0.9 43.8 32.6 33.0 109.4 29.2
Hybrid Conf. 90 3.3 63.0 42.7 39.8 145.5 72.6
Amine Process 90 8.2 77.5 45.5 44.9 (48.6) 167.9 (171.6) 106.3 (112.0)
Membrane 90 2.0 – 2.1 50.2 – 51.4 37.3 – 38.0 38.4 – 39.2 125.9 – 128.6 42.0 – 47.0
149
Chapter 5 Process and Cost Analysis of a Biomass Power
Plant

In this chapter, a novel concept is presented to capture CO2 from a large-scale


(>100 MWe) dedicated biomass-fired power plant by using CaO as the CO2 sorbent.
Biomass is burnt in a circulating fluidized bed at sufficiently low temperature to
allow in-situ CO2 capture at atmospheric pressure. A detailed process analysis is
carried out to reveal process performance and economics of the proposed power
plant which is compared against biomass-air-fired and biomass-oxy-fired power
plants. A heat exchanger network is designed using a pinch analysis aimed at the
recovery of the maximum amount of excess heat from high temperature gas and solid
streams in the plant while the recovered heat is transferred into a subcritical steam
cycle for power generation. The entire process simulation also includes a CO2
purification and compression unit that allows reaching more than 95 mol% CO2
purity. Work presented in this chapter is the outcome of my visit to Instituto
Nacional del Carbón (INCAR, Spain) where the part of work was carried out and has
resulted in one published article as detailed in Appendix E. Prof. Carlos Abanades
and Dr. Monica Alonso are co-authors of the publication and contributed with their
guidance and expertise.

5.1 Background

The use of sustainably-grown biomass as a renewable energy source in


combustion systems can be defined as a potential way to reduce CO2 emissions, as
this is associated with zero net CO2 emissions (Faaij et al., 1998). In Europe, it is
expected that the targets of 20% reduction in CO2 emissions by 2020 and increase in
the share of renewable energy to 20% can be partially achieved by replacing coal
with biomass in coal fired power generation systems (Maciejewska et al., 2006;
Böhringer et al., 2009).

Large scale (≥100 MWe) dedicated biomass-power plants have not been deployed
in the past for critical economic reasons linked to the limited local availability of
biomass sources, low energy density, higher moisture content and variability of fuel

150
characteristics. Most of the proposed solutions to introduce biomass in the power
generation sector have been aimed at co-firing biomass with coal in existing power
plants (Dai et al., 2008; Demirbas et al., 2009). It is generally agreed that co-firing up
to 20% biomass with coal is technically feasible with relatively modest modifications
of the existing system (Demirbas, 2003). However, further increases in biomass
content generate problems. For example high mixing ratios can create more
difficulties to fuel preparation and milling stages, boiler capacity and performance as
well as the utilization of ash, which reveals the requirement of a dedicated biomass
infrastructure (Maciejewska et al., 2006). In addition to a reduction of CO2
emissions, co-firing of biomass with high-sulphur coal can provide a reduction in
NOx and SOx emissions (Baxter, 2005). In the EU many countries have established
green certificate incentives for biomass use and this has led to an international
market of biomass feedstock for power generation (Schaeffer et al., 2013).

The latest developments in CFB technology are enabling a niche market in the
power sector for biomass combustion and gasification plants with net power
generation capacities between 100 and 250 MWe in standalone configurations or
linked to larger fossil plants (Stevens, 2001; IEAGHG, 2009). In particular, CFB
combustion systems for biomass have been defined as an ideal technology to be
deployed at large scale power generation from biomass co-firing owing to inherently
low emissions, high availability and superior fuel flexibility with high system
efficiencies (Hotta, 2010). The current CFB technology of dedicated biomass firing
is available up to 300 MWe and that for co-firing 50% biomass with coal can reach
600 MWe, while the maximum steam conditions for clean biomass firing can
approach 570°C and 180 bar, and for 50% biomass 585°C and 290 bar (Jäntti and
Nuortimo, 2012). The Foster Wheeler large scale utility at 300 MWe using an
advanced bio CFB boiler generates steam at 568/566°C and 179/43.6 bar and this
concept is market-ready (Jäntti and Nuortimo, 2012).

It should be noted that with current technology only subcritical steam cycles are
available for dedicated biomass-fired power plants while the supercritical steam
cycle conditions are only feasible for coal co-firing because of the high furnace tube
material temperatures and the requirement of tube materials to protect against

151
corrosion (Jäntti and Nuortimo, 2012). This limitation is linked, among other factors,
to the need to moderate combustion temperatures in the biomass combustor to
minimize risks of bed agglomeration, fouling and corrosion of the convective
surfaces (Leckner, 1998; Werther et al., 2000; Khan et al., 2009). Vaporized biomass
alkalis react with flue gas constituents and fly ash, producing low-melting species.
The potential of agglomeration, fouling and corrosion tendencies is linked to the
chlorine and sulphur contents and alkali composition so that the risks are more severe
for straw than for wood chips, sawdust and husk (rice) (Hiltunen et al., 2008; Jäntti
and Nuortimo, 2012).

5.2 Application of CO2 Capture to Biomass-fired Systems

The application of CO2 capture and permanent geological storage to biomass fired
systems leads to processes with negative emissions of CO2 (Ishitani and Johansson,
1996; Obersteiner et al., 2001; Rhodes and Keith, 2008). In theory all the CO2
capture technologies (pre-combustion, post-combustion and oxy-combustion) are
applicable for this purpose. The conceptual integration of a Ca-looping process for
in-situ CO2 capture from a dedicated biomass-fired power plant was proposed in a
previous study and is represented in Figure 5-1 (Abanades et al., 2011a). In the
combustor-carbonator reactor, biomass is fired with air and the CO2 generated in this
reactor is captured by CaO flowing from a calciner according to the following
reaction.

C (from biomass) + O2 (from air) + CaO → CaCO3 + Heat (Rn. 5.1)

The main additional product of the combustion is water, while for air the presence of
79 mol% inert nitrogen is assumed. The experimental performance of the proposed
design has been investigated in a 30 kWth test facility (Alonso et al., 2011a) and
scaled up to a 300 kWth pilot plant (Alonso et al., 2013). Although the work is still in
progress to characterize and optimize the operation of the larger pilot plant, the
operation of the simultaneous combustion-carbonation step in this system was found
to be technically feasible from a reactor perspective as combustion of biomass at
around 700ºC is known to yield high combustion efficiencies.

152
Gas Flow
Solid Flow
Possible Heat Recovery Points

CO2-depleted
flue gas CO2-rich gas

CaCO3, CaO
Purge (CaO)
Combustor Oxy-fired
Carbonator CaO Calciner Make-up (CaCO3)

Biomass Air Biomass Oxygen from Fan


ASU

Figure 5-1 Basic schematic diagram of the in-situ Ca-looping plant (Abanades et al., 2011a).

The combustor temperature of 700ºC also allows for theoretical CO2 capture
efficiencies over 80% when sufficiently active CaO is present in the combustor-
carbonator reactor which has been validated experimentally (Abanades et al., 2011b;
Alonso et al., 2011a; Alonso et al., 2011b). Despite the potential importance of the
concept shown in Figure 5-1 as a future application of bio-energy with CO2 capture
and storage (BECCS) technologies, there has not been a comprehensive process
analysis of the system and only a rough analysis of the cost structure has been
published so far (Abanades et al. 2011a). Therefore, the objective in this chapter is to
evaluate the system of Figure 5-1 in sufficient detail to allow for a comprehensive
cost analysis of the proposed system against similar alternatives for biomass
combustion with and without CO2 capture. A number of promising configurations
have been developed and the results have been compared against dedicated biomass-
air-fired and biomass-oxy-fired power plants in terms of net power generation
efficiency, CO2 capture rate and cost of CO2 avoided.

5.3 Reference Power Plants

5.3.1 Biomass-air-fired Power Plant

The CFB biomass-air-fired power plant used as a reference in this work has been
simulated according to relevant published reports (DOE, 2003; IEAGHG,2009) and

153
its schematic diagram including the detailed steam cycle configuration (DOE, 2007)
is shown in Figures 5-2(a) and 5-2(b), respectively. The plant uses a subcritical steam
cycle with a single reheat (166.5 bar/565.6°C/565.6°C). The steam cycle stream
properties for the plant from the simulations are given in Appendix D along with a
simulation of the effect of higher moisture contents. The steam cycle simulation used
in this study is calibrated against the reported data detailed in Ahn et al. (2013). As
the current biomass combustion technology does not allow severe steam generation
conditions, only the subcritical steam cycle is investigated for the proposed dedicated
biomass-fired power plant configurations (DOE, 2003; McIlveen-Wright et al., 2011;
Jäntti and Nuortimo, 2012).

The design parameters for the biomass-air-fired plant are given in Table 5-1. The
biomass-air-fired plant is an exemplary 135 MWe net output power plant with 40.9%
net efficiency based on lower heating value (LHV) of the design biomass (Abanades
et al., 2011a). The design biomass, with an LHV value of 20.08 MJth/kg estimated by
the simulator, is burnt in 15% excess air to guarantee complete combustion. The total
thermal energy input from biomass combustion is 329.8 MWth. The contents of ash
and alkalis in the biomass are not considered as limiting factors; although, depending
on the quality of the biomass, there would be the requirement for some modifications
in the steam cycle design. In the schematic diagram shown in Figure 5-2(a), the
combustion air is passed through a fan to provide 0.2 bar to overcome the pressure
drop in the combustor. The flue gases from the convection pass heat exchangers are
further cooled down to 135°C by pre-heating the inlet combustion air, and then flow
into a dust collector where particulates such as fly ashes are removed (DOE, 2003).
The CFB biomass-air-fired power plant includes an infiltration air flow rate
estimated to be 2 wt% of total gas flow rate from the combustor (DOE, 2003; DOE,
2007). The full process has been simulated in UniSim Design and the outputs of the
simulation are summarized in Table 5-1. The results are in close agreement with
those reported in the by the IEAGHG (2009) and provide a base for comparison with
simulations of the biomass-oxy-fired and in-situ Ca-looping concepts described in
the next paragraphs.

154
(a)

Gas Flow

Solid Flow
Convection
Pass Heat
Combustor Exchangers
Infiltration Air

Biomass Air Heater Air

Air fans

Ash
Dust Collector
Stack

(b)

Hot Reheat 1 14

5
Superheat 2 12
4 13

3
HP IP
CFB LP Turbine Generator
Turbine Turbine
Combustion 9 Boiler Feed
System 6 Pump Turbine
10
Drives
11 19
8 15 20
Superheater

21
22
Condenser
Economiser
Condensate
23
Pumps
7

Reheater 18

24
17
Cold Reheat
38
16
FWH 4 FWH 3 FWH 2 FWH 1
FWH 6 FWH 5 Deaerator
35 34 28 27 26 25

Feedwater
32 29 30 31
33
36 37

Boiler Feed
Pumps

Figure 5-2 Simplified schematic diagrams of the (a) biomass-air-fired power plant (DOE, 2003) and
(b) integrated subcritical steam cycle (IEAGHG, 2009). (HP: High Pressure; IP: Intermediate
Pressure; LP: Low Pressure; FWH: Feedwater Heater)

155
Table 5-1 Design specifications and performance summaries of the dedicated biomass-air-fired and biomass-oxy-fired power plants.

Biomass-air-fired Power Plant Biomass-oxy-fired Power Plant


Fuel Biomass Biomass
Fuel Composition (wt %) 5.4% H, 39.9% O, 0.7% N, 5.4% Ash, 41.4% C, 5.4% H, 39.9% O, 0.7% N, 5.4% Ash, 41.4% C,
7.2% Moisture 7.2% Moisture
Fuel Flow Rate (ton/h) 59.14 59.14
Fuel Lower Heating Value, LHV (MJth/kg) 20.08 (simulated) 20.08 (simulated)
Heat of Combustion (MWth) 329.8 329.8
Air/Oxygen Flow Rate (ton/h) 343.3 71.0
Infiltration Air Flow Rate (ton/h) 8.0 4.8
Steam Cycle Conditions (bar/°C/°C) 166.5/565.6/565.6 166.5/565.6/565.6
Gross Power Generation (MWe)/Efficiency (%) 143.8 / 43.6 147.6 / 44.8
Auxiliary Power (MWe) 8.8 8.0
Air Separation Unit Power (MWe) - 16.4
CO2 Purification and Compression Power (MWe) - 11.5
Net Power Generation (MWe)/Efficiency (%) 135.0 / 40.9 111.7 / 33.9
Overall CO2 Capture Efficiency (%) - 93
CO2 purity (mol %) - 97
156
5.3.2 Biomass-oxy-fired Power Plant

In the dedicated CFB biomass-oxy-fired plant, the thermal energy input from
biomass combustion is set to be the same as that in the biomass-air-fired plant, i.e.
329.8 MWth. The plant parameters for the biomass-oxy-fired plant are also reported
in Table 5-1. The oxygen feed at 99.1 mol% purity from an ASU is mixed with the
part of CO2-rich gases leaving the convection pass heat exchangers and is fed to the
combustor. The oxygen concentration in this mixture is fixed at 40 mol% to limit
flame temperatures in the combustor but it has a potential to be even higher (70
mol%) according to the DOE report (2003). The ASU power requirement is set as
231 kWh per ton of oxygen at 1.2 bar (DOE, 2003), which is identical to that used in
Chapter 2. The circulated CO2-rich gases are initially passed through a fan with a
final pressure of 1.2 bar to overcome pressure drop along the combustor. The gas
flowrate into the combustor is determined by setting the oxygen molar fraction in the
CO2-rich gas stream to 3 mol% and complete combustion is assumed. The CO2-rich
gases from the convection pass heat exchangers that are not recirculated to the
combustor feed, are initially cooled down to 172°C by exchanging heat with the inlet
oxygen feed and then further cooled down to 80°C in the second feedwater heater
(FWH2) shown in Figure 5-2(b) (DOE, 2003; DOE, 2010). Air infiltration is also
included as for the biomass-air-fired plant.

5.3.3 Power Plant Auxiliaries

In the two dedicated biomass power plants described above, a biomass drying unit
is not included. The limits of NOx emissions for biomass combustion systems can be
met by firing the boiler with low furnace exit temperatures, which can be completely
eliminated in the oxy-combustion systems (DOE, 2003; IEAGHG, 2009). Therefore,
there is no need for an SCR unit. Moreover, because of negligible content of sulphur
in the design biomass used in this study, an external FGD unit or limestone injection
is not included. The flow of infiltration air in the biomass-oxy-fired plant reduces the
CO2 concentration in the gas streams and leads to the requirement for a CO2
purification unit in order to achieve the target purity of ≥95 mol%.

157
Purge Gas
17

HE

20
Compressed
CO2 Expander
19
Pump
18

9 15
8 16

Separator Separator
HE
1 3 HE 4 5 10 11 12
Compressor
Train
2 Refrigeration I Compressor Refrigeration II
7 6

Waste Water 14 13
Pump Pump

Figure 5-3 Schematic diagram of the cryogenic CO2 purification and compression unit (Xu et al.,
2012). (HE: Heat exchanger)

A CO2 purification and compression stage (final pressure of 150 bar) is added to
the biomass-oxy-fired plant and the Peng-Robinson equation of state (Honeywell,
2010) is used as the thermodynamic model to calculate the phase equilibrium and
fluid properties. While only a simple CO2 compression design is needed owing to the
absence of air infiltration assumption in the Ca-looping and Ca-Cu looping process
schemes evaluated in Chapters 2 and 3, it is clear that a more advanced gas
processing system will be required in reality if there are any air-leakages into the
oxy-fired combustor (or calciner).

The configuration of the CO2 cryogenic liquefaction and separation system


proposed by Xu et al. (2012) has been adapted to the present case. The reference
cryogenic system comprises a two-stage compression, a two-stage refrigeration, a
two-stage separation and an energy recovery component. The pressure drops in the
heat exchangers will result in a maximum additional power requirement of 0.5 MWe
and as a result are not considered further. The cryogenic process has been modified
slightly because of the different feed gas compositions compared to 80 mol% in Xu
et al. (2012). In this study, the CO2 molar fraction in the feed is only 49 mol%, which
becomes 87 mol% on a dry basis and differs from the 80 mol% of Xu et al. (2012).
By adjusting slightly the gas pressure to reach it is possible to reach the targeted CO2
purity. The corresponding schematic diagram is presented in Figure 5-3 and the
stream compositions are given in Appendix D.

158
The CO2-rich gas stream is initially compressed to 24 bar by a series of
compressors with intermediate cooling to 35°C. The compressed gas is further
cooled to −20°C initially by heat exchange with liquid CO2 and then to −35°C by a
refrigeration cycle. The fluid refrigerant R502 with a boiling point of −45.4°C is
chosen for the first stage of refrigeration because of its acceptable thermodynamic
performance (Xu et al., 2012). Liquid CO2 from the first separator is pumped to 80
bar. In the second stage, the remaining gases are first further compressed to 54 bar
and then cooled to −18°C by heat exchange with liquid CO2 and purge gas from the
second separation stage. The refrigerant, R502, is also used in the second stage to
reduce the gas temperature to −35°C. The minimum temperature difference in the
low temperature heat exchanger is more than 2.1°C. The heat released during the
expansion of exhaust stream (No 16) is used to cool down the feed gas in the
compressor train. Finally, the liquid CO2 from the both stages of the cryogenic
separation process is pumped to 150 bar. The adiabatic efficiency of compressors and
pumps has been assumed to be 80%. The overall CO2 capture efficiency in the
cryogenic separation system is 93% with CO2 purity of ~97 mol%. The average
performance of vapour compression cycle coefficient (COP) was taken as 1.36 for
the first and second stages of refrigeration in order to estimate power use (Xu et al.,
2012).

5.3.4 Results and Discussion

The gross power generation capacity, presented in Table 5-1, is estimated to be


143.8 and 147.6 MWe for the biomass-air-fired and biomass-oxy-fired plants,
respectively. The difference in these values can be linked to the preheating
requirement of nitrogen in the air as well as lower level of heat recovery from the
flue gas stream which is fed to the stack at 135°C, compared to 80°C of the CO2 rich
stream in the biomass-oxy-fired plant. The power requirements associated with
auxiliaries (8.0 MWe), ASU (16.4 MWe) and CO2 purification and compression (11.5
MWe) in the biomass-oxy-fired plant results in a reduction of the net power output
down to 111.7 MWe with 33.9% LHV-based efficiency. The overall capture
efficiency and CO2 purity of the biomass-oxy-fired plant are directly related to the

159
efficiency in the CO2 purification and compression stage, which are 93% and 97
mol%, respectively.

5.4 In-situ Ca-looping Power Plant

5.4.1 Combustor-Carbonator Model

The key issue in the combustor-carbonator design is the selection of the operating
temperature since this is based on the equilibrium of CO2 over CaO and the
simultaneous combustion of biomass (Abanades et al., 2011a). It was concluded that
a reactor temperature of 700°C allows sufficiently high CO2 capture and combustion
efficiencies. At this temperature tar formation and the associated emissions of
hydrocarbons may be an issue for further research if the technology evolves towards
larger scale demonstration. However, considering the catalytic effect of CaO
particles on tar cracking reported by several authors in similar biomass gasification
processes in fluidized beds (Koppatz et al., 2009; Soukup et al., 2009) together with
the oxidizing nature of the combustor-carbonator, only minor traces of these
contaminants should be expected in the proposed systems and no cost penalty has
been allocated to treat them.

There are two heat sources in the combustor-carbonator which are the biomass
combustion and the exothermic carbonation reaction. Thus, a heat recovery system is
necessary to keep the reactor at the desired temperature. As for the previous designs
full conversion of the fuel is assumed. However, the detailed description of the in
situ Ca-looping plant requires additional assumptions on the level of conversion of
CaO to CaCO3 and the required level of solid circulation between the reactors. Even
though the rigorous carbonator model is employed in Chapters 2 and 3 for the
prediction of carbonation efficiency, the model is not applicable for this system
where simultaneous combustion and carbonation take place. Therefore, in this
chapter, the efficiency of carbonation is estimated by setting the conversion level of
CaO to CaCO3 (Xave) at 0.1 and the maximum average conversion rate of CaO
(Xmax,ave) at 0.15 considering that the sorbent is derived from a natural limestone
(Grasa and Abanades, 2006). The value of 0.05 (Xactive= Xmax,ave− Xave) as active
fraction of CaO particles is assumed to be enough to capture 80% of CO2 generated

160
in the carbonator which is in an agreement with the previous experimental work
where this was demonstrated (Alonso et al., 2011). Since the sorbent loses its
absorption capacity through cyclic carbonation and calcination cycles, part of the
sorbent is replaced with make-up sorbent that is assumed to be 100% CaCO3.

The quantity of purge flow rate has been calculated by using (Rodriguez et al.,
2010):

𝑎1 𝑓12 𝐹0 𝑎2 𝑓22 𝐹0
𝑋𝑚𝑎𝑥,𝑎𝑣𝑒 = + +𝑏 (Eq. 5.1)
𝐹0 𝐹𝑅 𝑓𝑐𝑎𝑟𝑏 (1−𝑓1 ) 𝐹0 𝐹𝑅 𝑓𝑐𝑎𝑟𝑏 (1−𝑓2 )

where a1, a2, b, f1 and f2 are decay constants obtained from experimental data
available for the CaO-based sorbent produced from natural limestone and taken as
0.1045, 0.7786, 0.07709, 0.9822 and 0.7905 respectively (Rodriguez et al., 2010).
fcarb refers to Xave/Xmax,ave ratio. With the known CO2 molar flow rate (FCO2), FR can
be calculated from the following equation where ECO2 refers to the carbonation
efficiency.

𝐹𝑅 𝑋𝑎𝑣𝑒
𝐸𝐶𝑂2 = (Eq. 5.2)
𝐹𝐶𝑂2

5.4.2 Assumptions to Build Mass and Energy Balances

The temperature of the calciner is set to 890°C which is 15°C higher than the
equilibrium temperature (Garcia-Labiano et al., 2002) and complete calcination is
assumed. The heat requirement mainly for the endothermic calcination reaction is
provided by combustion of biomass with oxygen from an ASU. A biomass-drying
unit is not included for consistency with the previous cases and also because in this
case it leads to a reduced CO2 partial pressure in the calciner, which allows to
operate the calciner at lower temperatures. Furthermore, it is shown that steam
addition into the calciner may potentially improve the cyclic stability and
performance of CaO-based sorbents (Rong et al., 2013). As in the previous oxy-fired
reference case, the purity of oxygen fed to the calciner is assumed to be 99.1 mol% at
1.2 bar and part of the exiting CO2-rich gases are circulated to lower oxygen
concentration in the feed gas to prevent high flame temperatures inside the calciner.
A circulation fan is required on the circulated CO2-rich gases to overcome pressure

161
drop in the calciner. As before, the inlet oxygen concentration was set equal to 40
mol%. The combustion air flowing into the combustor-carbonator is initially passed
through a fan to boost of the gas stream pressure by 0.2 bar to overcome the pressure
drop.

The overall mass and energy balances, presented in Table 5-2 for the schematic
diagram in Figure 5-4, have been conducted to achieve 80% CO2 capture efficiency
in the combustor-carbonator and 100% calcination efficiency. The resulting sorbent
to feed ratio (FR/FCO2) and purge rate ratio (F0/FCO2) are estimated to be 8 and 0.07,
respectively. The infiltration air flow rate into both reactors has been estimated to be
2% of total gas flow rate as for the previous cases. It has been assumed that 80% of
ashes from biomass combustion leave the system as fly ash. Moreover, 10% of
make-up (CaCO3) escapes the system as CaO in the fly ash because of sorbent
attrition. This portion of CaO is linked to a small additional energy penalty due to the
energy required for its calcination. It should be noted here that the assumptions of fly
ash and CaO attrition that improve the accuracy of process simulations are only
occupied in this chapter whereas they are not taken into consideration in Chapters 2
and 3. The total thermal energy input for the in-situ Ca-looping plant, which is the
sum of that for the combustor-carbonator and that for the calciner, is set to 329.8
MWth as before. Almost 64% of this energy needs to be supplied to the combustor-
carbonator while the rest is provided to the calciner.

5.4.3 Heat Recovery Steam Cycle

An important design issue for the in-situ Ca-looping plant is heat recovery from
high temperature gas and solid streams. There are five different heat recovery
locations on the diagram shown in Figure 5-1: combustor-carbonator; CO2-depleted
flue gas stream leaving the carbonator; CO2-rich gas from the calciner; solid purge
stream; and solid stream from the calciner to the combustor-carbonator. The last is
important since it reduces the quantity of heat that needs to be recovered in the
combustor-carbonator as well as the required heat transfer area in this reactor.

162
CO2-depleted
flue gases
CO2-rich gas
9
24
Bag
Fly ash
Filter
Q6 (+) Hot Reheat
23 Q4 (-) Superheat T7
Fly ash + CaO ESP
Q3 (+) T6
Infiltration Air HP
7 Infiltration Air
8 Turbine
18
0.83
Q5 (+) Q5 (-)
Q2 (+) T5
6 17
10 Q7 (+)
CaCO3, CaO Purge (CaO)
0.84 Q2 (-)
Q1 (+)
Combustor Oxy-fired 20 21
T4
Carbonator CaO Calciner Make-up (CaCO3)
T9

12 11 22 0.74 Q1 (-)
Q4 (+) 16
5 T3
15 19 0.26 Q1 (-)
Biomass Biomass
Circulation Fan
13 0.16 Q2 (-)
1 Q3 (-)
4 0.30 Q6 (-) Cold Reheat T2
14 T8
Oxygen from FWH 6
0.70 Q6 (-) ASU
3 0.17 Q5 (-) Feedwater
T1

Air

Figure 5-4 Detailed presentation of the heat recovery from the in-situ Ca-looping plant and its
integration with the reference steam cycle shown in Figure 5-2(b). The flow rate of steam in the steam
cycle is 400 t/h.

A heat exchanger network design is necessary to recover the maximum amount of


heat from this process, which can be then transferred to the steam cycle for power
generation. Even though a predefined gross efficiency has been employed to estimate
power generation capacity from high temperature sources in Chapters 2 and 3, the
steam cycle configuration described for the reference systems (Figure 5-2(b)) has
been modified in this chapter in order to recover heat from the recovery points
summarized above for the in-situ Ca-looping plant. The detailed integration of steam
cycle with the in-situ Ca-looping plant is shown in the right inset in Figure 5-4 and
corresponding stream conditions are presented in Table 5-2. The pinch methodology
is applied to estimate minimum energy requirement (MER) as well as to develop a
heat exchanger network design. The objective of pinch method is the reduction of
external energy requirements (heating or cooling) by maximizing heat transfer
between hot and cold streams (Linnhoff et al., 1978). A minimum temperature
difference of 20°C in the heat exchangers is assumed (Lara et al., 2013), while the
heat exchangers in the base steam cycle design have a minimum temperature
difference of 5°C (DOE, 2007).

163
Table 5-2 The stream properties and compositions for the in-situ Ca-looping plant presented in Figure 5-4.

Stream No 1 2 3 4 5 6 7 8 9 10 11 12
T (°C) 15 15 34.2 99 294 700 320 15 135 700 890 500
m (ton/h) 37.7 219.1 219.1 219.1 219.1 210.6 210.6 4.2 213.2 700.8 654.6 654.6
Composition (wt %)
CO2 - - - - - 5.4 5.4 - 5.4 - - -
O2 - 23.5 23.5 23.5 23.5 4.1 4.1 23.5 4.5 - - -
N2 - 76.5 76.5 76.5 76.5 79.7 79.7 76.5 80.3 - - -
CaO - - - - - - - - - 74.9 89.1 89.1
CaCO3 - - - - - - - - - 14.9 - -
C 41.4 - - - - - - - - - - -
H 5.4 - - - - - - - - - - -
S - - - - - - - - - - - -
O 39.9 - - - - - - - - - - -
N 0.7 - - - - - - - - - - -
H2O 7.2 - - - - 10.0 10.0 - 9.8 - - -
Ash 5.4 - - - - 0.8 0.8 - - 10.2 10.9 10.9
LHV (MJth/kg) 20.08 - - - - - - - - - - -
Total Stream Enthalpy (GJth/h) −71.23 −2.3 −2.0 16.6 61.1 −236.9 −335.0 −0.04 −357.0 −7757.0 −7076.0 −7322.0
164
Stream No 13 14 15 16 17 18 19 20 21 22 23 24
T (°C) 15 15 248 279.3 890 15 294.0 890 150 15 267.4 80
m (ton/h) 21.4 26.8 26.8 76.7 148.8 3.0 49.9 5.9 5.9 10.4 100.4 100.4
Composition (wt %)
CO2 - - - 53.7 83.3 - 82.6 - - - 82.6 82.6
O2 - 99.2 99.2 36.6 2.6 23.5 3.0 - - - 3.0 3.0
N2 - 0.8 0.8 2.0 1.1 76.5 2.6 - - - 2.6 2.6
CaO - - - - 0.4 - - 89.1 89.1 - - -
CaCO3 - - - - - - - - - 100 - -
C 41.4 - - - - - - - - - - -
H 5.4 - - - - - - - - - - -
S - - - - - - - - - - - -
O 39.9 - - - - - - - - - - -
N 0.7 - - - - - - - - - - -
H2O 7.2 - - 7.7 12.0 - 11.8 - - - 11.8 11.8
Ash 5.4 - - - 0.6 - - 10.9 10.9 - - -
LHV (MJth/kg) 20.08 - - - - - - - - - - -
Total Stream Enthalpy (GJth/h) −40.4 −0.25 5.6 −427.6 −1211.0 −0.03 −433.3 −63.7 −67.7 −125.7 −874.7 −895.2

Stream No T1 T2 T3 T4 T5 T6 T7 T8 T9
T (°C) 251.5 262.0 270.0 361.7 365.1 391.4 565.6 365.9 565.6
165
The pinch analysis specifications are summarized in Table 5-3 including all heat
duties. The target temperatures for the hot streams as well as the cold ones in the
steam cycle can be predefined whereas those for the combustion air and oxygen
streams should be calculated according to the energy balance after the steam cycle
integration. To reach maximum heat recovery with high power generation efficiency
the two primary target cold streams are feedwater at ~252°C and cold reheat at
~366°C in the reference steam cycle presented in Figure 5-2(b). The hot streams can
be cooled down to only 272°C, because of the minimum temperature difference of
20°C, when only those two streams are targeted. Further heat recovery from these
streams can be made either to the combustion air and oxygen streams or to the
feedwater heaters (FWH 1-6) in the steam cycle. In this study, the excess heat from
the CO2-rich gas and CO2-depleted gas streams are initially transferred to the
feedwater stream as shown in Figure 5-4 with final gas temperatures of 272°C and
320°C, respectively. Since the aim is to generate more steam at high temperature, the
remaining excess heat from those gas streams is transferred to the combustion air and
oxygen streams.

The feedwater is then further heated to the desired steam conditions by additional
heat from the combustor-carbonator and circulated solids at high temperature, while
the remaining heat from the combustor-carbonator is then sufficient to heat up the
cold reheat as suggested in Figure 5-4. Once this heat exchanger network
arrangement is defined, the flow rate of steam in the steam cycle is adjusted to
minimize the heat requirement of hot and cold utilities. With a steam flow rate of 400
ton/h, the MER is estimated to be almost zero, leading to the maximum theoretical
heat recovery from the in-situ Ca-looping plant to the steam cycle. For this threshold
problem, the pinch point is located at 870 - 890°C and the corresponding Grand
Composite Curve (GCC) is shown in Figure 5-5. It should be mentioned that
potential heat transfer from the purge stream to the steam cycle is not considered
because of its low heat duty. There would be a marginal increase in the required total
heat exchange area in the in-situ Ca-looping configuration as the combustor is
operated at a lower temperature than commercial boilers (~850°C) but this issue is
not considered in the current level of development.

166
Table 5-3 Pinch analysis specifications for the in-situ Ca-looping plant.

Supply Target Heat Duty Stream


Heat Recovery Point Temperature (°C) Temperature (°C) (MWth) Stream Type
Combustor-Carbonator (C-C) 700.0 700.0 183.8 +Q1 Hot
CO2-depleted Gas 700.0 135.0 39.6 +(Q2+Q3) Hot
Solids from Calciner to C-C 890.0 500.0 68.3 +Q4 Hot
CO2-rich Gas (from calciner) 890.0 272.0 32.8 +Q5 Hot
CO2-rich Gas (to compression) 267.4 80.0 5.7 +Q6 Hot
Purge from Calciner 890.0 150.0 1.1 +Q7 Hot
Combustion Air into C-C 34.2 294.0 16.4 −(Q3+0.71 Q6) Cold
Oxygen into Calciner 15.0 248.0 1.7 −(0.30 Q6) Cold
Cold Reheat 365.9 565.6 48.0 −(0.26 Q1) Cold
Feed-water 251.5 565.6 264.1 −(0.74 Q1+Q2+Q4+Q5) Cold
167
1000

800

Shifted Temperature (°C)


600

400

200

0
0 50000 100000 150000 200000 250000 300000
Net Heat Flow (kW)

Figure 5-5 The grand composite curve prepared according to the specifications given in Table 5-3.

5.4.4 Results and Discussion

Table 5-4 shows the energy balances around the proposed power plant schemes.
The difference between the total enthalpies of the inlet and outlet streams is resulting
from the amount of heat transferred into the steam cycle and the energy consumed by
the fans. The enthalpy of the biomass feed is same in the all cases due to the constant
flow rate of the biomass whereas the air/oxygen flow rate including the infiltration
air differs between cases. The major term having priority on the gross power
generation efficiency estimations is the total heat transfer into the steam cycle. This
value is the greatest for the biomass-oxy-fired plant. The expectation would be that
the in-situ Ca-looping plant would be second as it includes both air-fired and oxy-
fired combustors. However, due to the calcination energy requirement (5.2 MWth)
and heat losses from the purge stream (1.1 MWth); it turns out that the heat flow from
the in-situ Ca-looping process into the steam cycle is slightly lower than that for the
biomass-air-fired plant. The heat spent for calcination can be discounted if the purge
stream is used in a cement plant but this option has not been considered here. There
are two components in the heat flow to the steam cycle from the biomass-oxy-fired
plant: the majority of heat flow (321 MWth) is used to generate steam and on the cold

168
reheat stream while the remaining 4.0 MWth reduce the heat duty on the second feed
water heater. The other two configurations only use energy for steam generation and
for the cold reheat.

The key design parameters and outcomes of the in-situ Ca-looping plant
simulation are summarized in Table 5-5. The CO2 purification unit is essential to
increase the CO2 purity to the target purity of more than 95 mol%. The cryogenic
CO2 purification-compression process is very similar to the one described previously
and presented in Figure 5-3. The corresponding power requirement is reported in
Table 5-5. The final CO2 product is compressed to 150 bar as for the biomass-oxy-
fired plant. The thermal energy transferred into the steam cycle is consistent with the
case compiled in Table 5-4, and the gross power generation capacity of this plant is
143.0 MWe. This value reduces to 117.9 MWe (35.7% net efficiency) when
additional power losses linked to auxiliaries, ASU and CO2 purification-compression
are included. The CO2 capture efficiency obtained is 88% before the CO2
purification-compression stage and reduces to 84% when the efficiency of the
cryogenic separation is included. Overall, the net power generation efficiency of the
in-situ Ca-looping plant is higher than that for the biomass-oxy-fired plant but it
should be noted that the capture efficiency is higher for the latter.

Table 5-4 Energy balances of the proposed dedicated biomass power plants.

Biomass-air- Biomass-oxy- In-situ Ca-


fired fired looping
Power Plant Power Plant Power Plant
(MWth) (MWth) (MWth)
Enthalpy In (A)
Biomass −31.0 −31.0 −31.0
Air/Oxygen −1.0 −0.2 −0.7
Make-up - - −34.9
Enthalpy Out (B)
Fly/Bottom Ashes −12.1 −12.1 −11.5
Exit Gas Streams −331.9 −343.1 −347.9
Purge - - −17.7
Heat of Compression by Fans (C) 1.8 1.0 1.6
Heat into the Steam Cycle (D) 313.8 321.0 + 4.0 312.1
Net ((B−A)+(D−C)) 0 0 0

169
Table 5-5 The in-situ Ca-looping plant process specifications and performance summary.

In-situ Ca-looping Power Plant


Fuel Biomass
(see Table 5-1 for composition)
Heat of Combustion (MWth)
Combustor/Carbonator 210.4
Calciner 119.4
Air/Oxygen Flow Rate (ton/h) 219.1 / 26.8
Flue Gas Composition before CO2 Purification (mol %) 3.5% O2, 3.4% N2, 24.1% H2O,
69.0% CO2
Flue Gas Composition after CO2 Purification (mol %) 1.6% O2, 1.3% N2, 0.3% H2O,
96.8% CO2
Steam Cycle Conditions (bar/°C/°C) 166.5/565.6/565.6
Gross Power Generation (MWe)/Efficiency (%) 143.0 / 43.3
Auxiliary Power (MWe) 8.5
Air Separation Unit Power (MWe) 6.2
CO2 Purification and Compression Power (MWe) 10.4
Net Power Generation (MWe)/Efficiency (%) 117.9 / 35.7
CO2 Capture Efficiency before the Purification Stage (%) 88.0
Overall CO2 Capture Efficiency (%) 84.1

5.5 Economic Analysis

The aim of this section is to provide a simple and transparent economic analysis
for the estimations of costs of electricity (COE) and CO2 avoided associated with the
biomass-fired power plants discussed in this chapter. The following equation has
been used to estimate COE (€/kWhe) (Abanades et al., 2007; Abanades et al., 2011a).
The specifications of the economic analysis have been summarized in Table 5-6.

𝑇𝐶𝑅×𝐹𝐶𝐹+𝐹𝑂𝑀 𝐹𝐶
𝐶𝑂𝐸 = + 𝑉𝑂𝑀 + (Eq. 5.3)
𝐶𝐹×8760 𝜂

where TCR (€/kWe) is the total capital requirement and FCF is the fixed charge
factor which is assumed to be 0.1 (Abanades et al., 2011a). FOM refers to the fixed
operating and maintenance costs and is taken as 3.70% of the TCR (SETIS, 2013).
CF is described as the capacity factor and set to 90% for all cases. For the sake of
simplicity, the variable operating and maintenance costs, VOM, is fixed at 0.01
€/kWhe (SETIS, 2013). FC (€/kWhth) is the fuel cost and selected depending the type
of fuel (either coal or biomass) and η (kWhe/kWhth) is the net electrical efficiency.

170
Table 5-6 The main specifications used for the economic analysis.

Value
Reference Biomass-air-fired CFB Power Plant*
Capacity (MWth)** 654.0
Capital Cost, TCR (×1000€)
Solid Storage and Handling 49,000
Boiler Island 204,000
Power Island 60,800
Utilities & Offsites 56,400
Base Year 2009
Fixed Fraction Cost, FOM (% TCR) 3.70
Capacity Factor, CF (%) 90
Fixed Charge Factor, FCF 0.1
Variable Cost, VOM (€/kWhe) 0.01
Coal Cost (€/kWhth) 0.01
Biomass Cost (€/kWhth) 0.03
ETS (€/ton CO2) 14
Green Certificate (€/MWhe) 50
Scale Factor 0.6
* Data from IEAGHG (2009).
** Based on LHV of the fuel.

The TCR of the reference biomass-air fired plant has been taken from the study on
biomass plants with CCS (IEAGHG, 2009) and is reported in Table 5-6 together with
its thermal energy capacity and base year. The TCR of this plant comprises of the
costs of solid storage and handling unit; boiler island; power island and utilities. The
cost data available for the biomass-oxy-fired case is scarce in the literature.
However, the similarities of biomass-fired systems with existing coal-fired CFB
power systems can be taken into consideration. According to DOE (2003), the total
TCR of the boiler island in a coal-air-fired plant is 60% higher than that for a coal-
oxy-fired plant at same thermal energy input, which can be linked to the volume of
the boilers. Thus, it is possible to estimate the cost of the boiler island from the
values given for the reference biomass-air-fired plant. In addition, the TCRs of the
ASU and CO2 purification-compression unit in the biomass-oxy-fired plant have
been calculated based on their power requirements compared to the reference units
given in Table 4-2 (DOE, 2003).

For all proposed cases, TCR calculations have been based on the same thermal
energy input, 329.8 MWth. The six-tenths rule and the M&S index detailed in
Chapter 4 are used to approximate TCR of a plant or a unit. Calculations for the
biomass-air-fired and biomass-oxy-fired plants are straight forward. However, more

171
attention has to be given for the case of the in-situ Ca-looping plant. This system
contains an air-fired CFB combustor and an oxy-fired CFB combustor at smaller
scale. Also, the ASU and CO2 purification-compression unit existing in this system
are smaller compared to these in the biomass-oxy-fired plant. As indicated,
approximately 64% of the total thermal energy input is supplied into the combustor-
carbonator as an air-fired CFB combustor. The rest of energy is used in the calciner,
as an oxy-fired CFB combustor. Thus, the air- and oxy-fired combustors in the in-
situ Ca-looping plant have been scaled based on the individual thermal energy input
into these reactors whereas the TCRs of the solid storage and handling unit, power
island and utilities have been estimated based on the total thermal energy input as
these units are shared by both combustors. As mentioned previously, the heat
exchange area requirement in the combustor-carbonator should be higher than the
reference CFB boiler operated at higher temperature. However, the cost increase
linked to variation in the heat transfer area is not considered in this study and a more
detailed study is required to calculate the heat transfer area in both cases.

The calculated TCR results are shown in Table 5-7, together with COE and cost
of CO2 avoided estimations. To compare the economic performance of the biomass
power plants against a CFB coal-air-fired power plant, the specific CO2 emission and
the TCR of a reference CFB coal-air-fired plant are taken from the DOE report
(2003). To have a more accurate estimation, the TCR of the reference coal-fired plant
is adjusted by the six-tenths rule and the M&S index based on the same thermal
energy input given for the biomass power plants. The estimated TCRs of the coal-
and biomass-fired CFB power plants, presented in Table 5-7, are lower than those of
the capture options since there are no ASU and CO2 purification-compression unit.
One could expect a lower TCR for the in-situ Ca-looping plant since the capacity of
the ASU for this system is lower compared to the biomass-oxy-fired plant. The
calculated TCRs are very close and this can be explained by the difference in the
capital costs of air- and oxy-fired CFB boilers as well as the use of the scale factor in
the calculations. While the estimated COE is the lowest for the coal-air-fired plant,
the addition of capture units increases the cost further with a maximum of 139
€/MWhe estimated for the biomass-oxy-fired plant.

172
Table 5-7 Comparison of the dedicated biomass-fired power plants in terms of costs of electricity (COE) and CO2 avoided.
Coal-fired CFB Biomass-air-fired Biomass-oxy-fired In-situ Ca-looping
Specific CO2 Emission (kg CO2/MWhe) 998 665 56 127
CO2 emission factor (kg CO2/MWhe) 998 0 −744 −672
TCR (€/kWe) 1660 1903 3012 2816
COE (€/MWhe) 60 109 139 132
COE inc ETS (€/MWhe) 74 109 128 122
COE inc ETS and GC (€/MWhe) 74 59 78 72
Cost of CO2 Avoided (€/ton CO2 avoided) - 49 45 43
Cost of CO2 Avoided inc ETS (€/ton CO2 avoided) - 35 31 29
Cost of CO2 Avoided inc ETS and GC (€/ton CO2 avoided) - −15.3 2.4 −1.1
Case A
TCR (€/kWe) - - 2887 2697
Cost of CO2 Avoided* (€/ton CO2 avoided) - - 1.5 −2.1
Case B
TCR (€/kWe) - 1759 2628 2084
Cost of CO2 Avoided* (€/ton CO2 avoided) - −17.2 −0.5 −6.9
* Including ETS and green certificate (GC) incentives.
173
The additional benefits from the ETS as in Chapter 4 and green certificates
(IEAGHG, 2009) have been included in the economic analysis. The reference ETS
price is fixed at 14 €/ton CO2. The green certificates are used to include the
environmental value of renewable energy and set to 50 €/MWhe. With the
introduction of the ETS and green certificate incentives, the COE estimations for the
capture cases, biomass-oxy-fired and in-situ Ca-looping plants, drastically decrease
in comparison with the coal-fired plant. The COE of the in-situ Ca-looping plant
reduces to 72 €/MWhe and it becomes even lower for the biomass-air-fired plant (59
€/MWhe).

To calculate cost of CO2 avoided (€/ton CO2 avoided) for the biomass combustion
systems (bio), the following general equation can be used.

𝐶𝑂𝐸𝑏𝑖𝑜 −𝐶𝑂𝐸𝑐𝑜𝑎𝑙−𝑓𝑖𝑟𝑒𝑑
𝐶𝑜𝑠𝑡 𝑜𝑓 𝐶𝑂2 𝑎𝑣𝑜𝑖𝑑𝑒𝑑 = (Eq. 5.4)
𝐶𝑂2 𝑒𝑚𝑖𝑠𝑠𝑖𝑜𝑛 𝑓𝑎𝑐𝑡𝑜𝑟𝑐𝑜𝑎𝑙−𝑓𝑖𝑟𝑒𝑑 −𝐶𝑂2 𝑒𝑚𝑖𝑠𝑠𝑖𝑜𝑛 𝑓𝑎𝑐𝑡𝑜𝑟𝑏𝑖𝑜

where the CO2 emission factorcoal-fired is equal to the specific CO2 emission in the coal
reference plant. However, the definition of the CO2 emission factor for the biomass
combustion systems must consider the neutral character assumed for the carbon from
biomass and the overall efficiency of CO2 captured (Ecapt). In the case of the biomass
air-fired plant, the emission factor is zero (IEAGHG, 2009), while this takes negative
values when CO2 capture is applied to the plant. The CO2 emission factors as well as
the specific CO2 emissions for all the systems considered in this work are also
summarized in Table 5-7. For the biomass-oxy-fired and in-situ Ca-looping plants
referred as ‘bio,cc’, the CO2 emission factor is calculated by:

𝐸𝑐𝑎𝑝𝑡
𝐶𝑂2 𝑒𝑚𝑖𝑠𝑠𝑖𝑜𝑛 𝑓𝑎𝑐𝑡𝑜𝑟𝑏𝑖𝑜,𝑐𝑐 = −𝑠𝑝𝑒𝑐𝑖𝑓𝑖𝑐 𝐶𝑂2 𝑒𝑚𝑖𝑠𝑠𝑖𝑜𝑛𝑠𝑏𝑖𝑜,𝑐𝑐 (Eq. 5.5)
�1−𝐸𝑐𝑎𝑝𝑡 �

The cost is the minimum for the in-situ Ca-looping process at 43 €/ton CO2
avoided, when excluding ETS and green certificate incentives whereas this decreases
to −1.1 €/ton CO2 avoided when the incentives are included. An even more attractive
negative avoided cost result is calculated for the biomass-air–fired plant mainly
because of the inclusion of green certificates. The cost of CO2 avoided results
calculated for the biomass-oxy-fired and in-situ Ca-looping plants are similar. While

174
net power generation efficiency of the in-situ Ca-looping plant is slightly higher than
that for the biomass-oxy-fired plant, the overall CO2 capture efficiency is greater for
the latter. Thus, the similarity can be explained by a trade-off between those values.

Two different additional approaches have been considered for TCR estimations of
the dedicated biomass-fired power plants when they are built next to other
combustion systems. The results of these cases can also be seen in Table 5-7. In Case
A, it is considered that biomass-oxy-fired and in-situ Ca-looping plants are able to
share their oxygen supply with another oxy-fired system located nearby. In this case,
it is assumed that the total oxygen requirement in the plants, a dedicated biomass-
fired power plant with CO2 capture and the other oxy-fired plant, equals that of the
reference ASU shown in Table 4-2. The capital cost of the ASU in the biomass-fired
plants can be then calculated from the flow of oxygen to these systems. In other
words, this is equivalent to eliminating the scale factor in TCR calculations for the
ASU. The total TCR estimations reduce to 2887 and 2697 €/kWe for the biomass-
oxy-fired and in-situ Ca-looping plants, respectively while the corresponding cost
including ETS and green certificate incentives decreases to −2.1 €/ton CO2 avoided
for the latter.

In more advanced process integrations (Case B in Table 5-7), the biomass-fired


plants can be coupled directly with another fossil-based power generation systems to
reduce CO2 emissions since biomass-firing is linked to zero CO2 emission, for
example parallel co-firing, where coal and biomass are burnt in separate boilers. In
this type of integration, the biomass-air-fired plant can share power island and
utilities costs with the other power plant while biomass-oxy-fired plant can also
benefit from a larger ASU and CO2 purification-compression unit if it is integrated
into an oxy-fired power plant. For the in-situ Ca-looping plant, in addition to above
benefits, the most advantageous integration can be achieved if this process is coupled
with another post-combustion Ca-looping unit that is used to capture CO2 from a
power plant. The availability of CaO from the calciner at larger scale will eliminate
the requirement of an additional calciner as well as its capital requirement. The
capital cost of a calciner in the in-situ Ca-looping plant can be estimated based on its
energy requirement over the total thermal energy input into the larger scale calciner.

175
For Case B calculations, similar to Case A, it is assumed that the total thermal energy
requirement in a biomass-fired plant and a fossil-based power system equals to that
for the reference biomass-air-fired plant given in Table 5-6. Further reduction in total
TCR can be achieved in Case B while the level of reduction is more severe for the in-
situ Ca-looping process because the units existing in this system are normally at
smaller scale compared to the other biomass-fired plants. The cost reduces down to
−17.2 €/ton CO2 avoided for the biomass-air-fired plant and −6.9 €/ton CO2 avoided
for the in-situ Ca-looping plant but the specific CO2 emission is much lower for the
latter. In summary, owing to the six-tenths rule, this exercise indicates that
substantial reductions in the avoided cost can be obtained if the small scale biomass
systems can use specific capital cost characteristic of much larger power generation
systems through integration. It is important to note here that direct co-firing in the
existing systems could also be an option, but there is a general trend in reducing the
legislative financial support to biomass co-firing and in practice it may be
economically more favourable to build a separate unit which uses only biomass. For
example in the UK biomass co-firing will attract only 0.5 Renewable Obligation
Certificates (ROCs) per MWh of electricity produced, while a dedicated biomass
plant will receive 1.4 ROCs for the same output (DECC, 2012b).

Finally, the sensitivities of the cost calculations with respect to some of the cost
parameters adopted above have been examined based on a ±25% variation (see
Figure 5-6). The CO2 avoided estimations including ETS and green certificate
incentives already shown in Table 5-7, excluding Cases A and B, are selected as
bases for the sensitivity analyses. The influence of the biomass cost is the largest,
especially on the biomass air-fired plant. The green certificate subsidy has a similar
influence, as it directly affects the effective fuel cost for the power plants. In cost
scenarios where the ETS price is sufficiently high the biomass oxy-fired system
becomes the preferred option. The in-situ Ca-looping technology is only slightly
more economical when both ETS and green certificate incentives are in place. The
impact of scale factor is milder in the sensitivity analyses compared to the advanced
cases (A and B) investigated above since it also changes the TCR of the reference
coal-air-fired plant.

176
(a)

40

[€/ton CO2 avoided]


20
Cost

-20

-40
Biomass Cost TCR ETS GC Scale Factor

(b)

40

20
[€/ton CO2 avoided]
Cost

-20

-40
Biomass Cost TCR ETS GC Scale factor

(c)

40
[€/ton CO2 avoided]

20
Cost

-20

-40
Biomass Cost TCR ETS GC Scale factor

Figure 5-6 Sensitivity analysis on cost parameters for the a) biomass-air-fired plant; b) biomass-oxy-
fired plant; c) in-situ Ca-looping plant. (TCR: Total Capital Requirement; ETS: European Trading
Scheme CO2 cost; GC: Green Certificates)

177
The option of a biomass air-fired plant clearly becomes the most competitive
option to avoid CO2 emissions at present because many countries have adopted a
policy of green certificates, while the ETS is very low. If this policy is sustained, this
could lead to a deployment of air-fired CFBC systems with no CO2 capture. If there
was later a policy to favour CCS systems (much higher cost of ETS credits), the
retrofitting of existing air-fired CFB biomass combustion plants to CCS could be
facilitated with the in situ Ca-looping system as this would allow not only the capture
of the CO2 generated in the oxy-fired calciner but also the capture of the CO2
generated in the existing air-fired plant. This added flexibility in the in-situ system
has not been incorporated in the cost analysis as it is too sensitive to uncertain
assumptions on the residual value of the air-fired plant. It is sufficient to conclude at
present that all three systems discussed in this work have a chance to be highly
competitive to avoid CO2 emissions depending on political incentives determining
the values of ETS or green certificates.

On the other hand, the current capacity of the dedicated biomass power plants in
the UK was reported to be 400 MWe, and that for co-firing is 255 MWe (DECC,
2014). It is estimated that around 1700 km2 of land to grow the fuel on will be
required if these plants are running with 90% capacity and using only energy crops
(DECC, 2014). Therefore, it is worth to consider that there would be limitations on
the availability of biomass if several large scale dedicated biomass power plants
become operational.

5.6 Concluding Remarks

The conceptual process integration of an in-situ Ca-looping process with a


dedicated biomass power plant allows effective removal of CO2 from biomass
combustion when the combustor-carbonator operates at around 700°C. Thanks to the
biomass-fired oxy-calciner, most of the CO2 from combustor-carbonator as well as
the entire fraction of CO2 from biomass combustion in this reactor can be removed at
high purity. This system, with a heat exchanger network design, is capable of
achieving a greater net power generation efficiency compared to a biomass-oxy-fired
plant. An overall CO2 capture efficiency of 84% is achieved by capturing 80% CO2

178
resulting from biomass combustion in the combustor-carbonator while 93 % capture
efficiency is estimated for the biomass-oxy-fired plant. The net power generation
efficiency decreases from 40.9% for the biomass-air-fired plant to 35.7% in the
proposed system mainly because of the requirements of the ASU and CO2
purification-compression sections.

A detailed cost analysis has been presented which allows the estimation of
electricity and cost of CO2 avoided with and without incentives, i.e. ETS carbon
price and green certificates. The lowest cost of electricity is obtained for a coal-fired
power plant without CO2 capture due to lower cost of coal compared to biomass.
With the inclusion of ETS and green certificate incentives (14 €/ton CO2 and 50
€/MWhe, respectively), the cost of electricity is found to be minimum for the
biomass-air-fired plant. The cost of CO2 avoided estimates for the biomass-oxy-fired
and in-situ Ca-looping plants are close but the gap widens with different process
integration possibilities. The sensitivity analyses on economical estimations reveal
that costs of biomass and revenue from green certificates have significant impacts on
cost of CO2 avoided calculations. Although several different large-scale dedicated
biomass power systems with CO2 capture have been studied in this chapter, this
approach may not be the optimal one for smaller scale biomass plants, which are
more common, and that further studies are needed to evaluate this.

179
Chapter 6 Conclusions and Recommendations for
Future Work

The objective of this dissertation is to conduct the techno-economic assessment of


carbon capture technologies to reduce CO2 emissions from cement and biomass-fired
power plants. The Ca-looping process has emerged as a leading option for this
purpose because of the use of low cost sorbent, the relatively small energy penalty,
and the possibility of using purged sorbent for clinker production. Various alternative
carbon capture technologies including membrane, amine, oxy-combustion and
indirect calcination processes have also been investigated, and the comparison of
these processes with the Ca-looping process has been conducted. Detailed process
flowsheets have been generated using Honeywell’s UniSim R400 to reveal process
performance and economics of the proposed carbon capture technologies. The most
significant conclusions and recommendations for future work are outlined in this
chapter.

6.1 Conclusions

6.1.1 Ca-looping Process for Cement Plants

A way of capturing CO2 from cement plants by using the Ca-looping process has
initially been investigated. The base cement process simulation implemented in this
study was proven to be reliable by comparing the estimated composition of clinker
and thermal energy requirement with those reported in the literature. The flue gas
stream leaving the 3rd preheater has been selected to be the optimal feed for the
carbonator because: (i) it has higher CO2 concentration and a lower total volumetric
flowrate; (ii) it does not have to be preheated; and (iii) it facilitates the design of a
heat exchanger network for heat recovery. The CO2-rich gas stream from the calciner
was directly sent to the CO2 compression unit rather than the preheater (Rodriguez et
al., 2012) as this may result significant reduction in CO2 purity if any air-leakage into
this unit cannot be prevented. A model for a carbonator, where hydrodynamics in fast
fluidisation, reaction kinetics as well as active fraction of the sorbent are taken into
account, has been incorporated into the cement process flowsheet. The Ca-looping

180
process can achieve over 90% CO2 avoidance with energy consumption ranges from
2.5 to 3.0 GJth/ton CO2 avoided depending on the F0/FCO2 ratio.

6.1.2 Ca-Cu Looping Process for Cement Plants

The potential of incorporating a CLC cycle into the Ca-looping process as an


alternative way to transfer oxygen into the calciner was carefully investigated. The
configuration with a solid route of carbonator → calciner (fuel reactor) → air reactor
→ carbonator has been selected to provide an opportunity for heat transfer between
the calciner and air reactor by hot solid circulation. To the best of our knowledge,
this was the first approach that addressed the need of heat transfer between these
reactors to prevent severe heat requirements in the proposed system. Up to 1 wt%
CuO from the purge was allowed in the cement raw meal for clinker production,
which corresponds to a F0/FCO2 ratio of 0.15. The incremental energy consumption
required by this system was estimated to be as low as 1.7 GJth/ton CO2 avoided,
which is at least 30% lower than that of the conventional Ca-looping process.

6.1.3 Alternative Carbon Capture Technologies for Cement Plants

A detailed process integration of the indirect calcination process into the base
cement plant has been analysed. The major advantage of the proposed configuration
is the use of surplus heat from high temperature gas streams for raw meal preheating
rather than power generation suggested in Rodriguez et al. (2011b). The incremental
energy consumption of this system was estimated to be 0.9 GJth/ton CO2 avoided at
56% CO2 avoidance. The energy consumption increases to 3.3 GJth/ton CO2 avoided
in the new hybrid configuration which includes an additional amine capture unit,
while that of the standalone amine process was calculated to be 8.2 GJth/ton of CO2
avoided. Furthermore, a dual stage membrane configuration comprised of a counter-
current module with a sweep, followed by a cross flow model has been implemented
for two feed gas locations, i.e. preheater gas or end-of-pipe gas. It was observed that
the membrane area was 48% higher for the latter case. The incremental energy
consumption for the membrane process was calculated to be as low as 2.0 GJth/ton
CO2 avoided at 90% CO2 avoidance, which is less than that of the equivalent Ca-
looping process.

181
6.1.4 Economic Analysis of the Carbon Capture Technologies Applied to
.Cement Plants

A method of economic analysis has been applied to evaluate the cost performance
of the proposed carbon capture technologies. The close similarity of the CFB reactors
existing in certain carbon capture systems with commercial CFB power plants were
taken into account for capital cost estimations. The LCOC for the base cement plant
increases from 97 €/ton cement to around 127 €/ton cement when the Ca-looping
process is incorporated and stays almost constant at different F0/FCO2 ratios.
Including the ETS price, the avoided cost was calculated to be around 41 €/ton CO2
for the Ca-looping process. The economic performance of the Ca-Cu looping process
is not competitive with the Ca-looping process at very high F0/FCO2 ratios because of
the severe variable cost associated with the purged CuO/Al2O3. The avoided cost
increases up to 157 €/ton CO2 for the Ca-Cu looping process at 0.15 F0/FCO2. It is
clear that separation and reuse of the purged CuO/Al2O3, depending on the stability
of the sorbent, are necessary to make this process more economical.

Amongst the various carbon capture technologies compared in this dissertation,


the minimum cost belongs to the standalone indirect calcination process with a value
of 29 €/ton CO2 avoided. This cost further increases to 73 €/ton CO2 avoided in the
hybrid configuration aiming at 90% CO2 avoidance, but it is around 30% lower
compared to that of the standalone amine process. In the case of the membrane
process, the LCOC increases to 125 €/ton cement for the preheater integration and
129 €/ton cement for the end-of-pipe integration. The cost of CO2 avoided for these
cases were calculated to be 42 €/ton CO2 and 47 €/ton CO2, respectively. It should be
highlighted that although the Ca-looping process requires greater energy
requirement, the final cost estimate for this process and membrane process are very
similar owing to the benefit of on-site power generation.

6.1.5 Dedicated Biomass Power Plant with in-situ Ca-looping Process

The first comprehensive process analysis of a large-scale dedicated biomass


power plant with in-situ CO2 capture using the Ca-looping process has been reported.
This system was compared against similar alternatives, biomass-air-fired and

182
biomass-oxy-fired power plants in terms of process and cost performances. A
subcritical steam cycle with a single reheat that is currently the only available
technology for biomass combustion has been included in the process configurations.
The net LHV-based power efficiency of 40.9% for the biomass-air-fired power plant
reduces to 33.9% for the biomass-oxy-fired power plant. The net efficiency of the in-
situ Ca-looping plant achieving 84% overall CO2 capture efficiency was calculated to
be 35.7%, representing an energy penalty of 5.2%. Furthermore, the COE estimate
increases from 109 €/MWe for the biomass-air-fired plant to 139 €/MWe and 132
€/MWe for the biomass-oxy-fired and in-situ Ca-looping plants, respectively. With
the inclusion of the ETS price and green certificate incentives, the COE reduces to 59
€/MWe and 72 €/MWe for the biomass-air-fired and the in-situ Ca-looping plants,
respectively. It reflects a negative cost of CO2 avoided estimate for the in-situ Ca-
looping plant with a value of -1.1 €/ton CO2 avoided. It was concluded that the
biomass-air-fired plant taking the advantage of green certificates is the preferable
option if the ETS price is very low, whereas the capture processes will be favourable
if the ETS price is much higher.

6.2 Recommendations for Future Work

Further work may be carried out based on the following recommendations.

• A detailed steam cycle configuration and a heat exchanger network design were
explored for heat recovery from the biomass power plant with in-situ Ca-looping
system detailed in Chapter 5. The same methodology can be adapted in Chapters
2 and 3 for the post-combustion Ca-looping and Ca-Cu looping processes to
improve the prediction of power generation efficiency in these systems.
• The use of more realistic assumptions regarding sorbent attrition and air-leakage
in the Ca-looping, Ca-Cu looping and indirect calcination processes presented in
Chapters 2 and 3 would improve the accuracy of current predictions. It is
expected that the sorbent attrition results an increase in calciner heat duty and
variable cost, while any air-leakage into the calciner will arise the requirement of
CO2 purification stage as presented in Chapter 5.

183
• Oxy-kiln would be a possible alternative to increase CO2 avoidance rate more
than 90% when it is coupled with an oxy-fired calciner if the future studies
approve the utilization of oxy-combustion in the kiln.
• If detailed calciner (fuel reactor) and air reactor modelling studies are conducted,
more accurate predictions can be obtained for the calcination, oxidation and
reduction efficiencies in these reactors.
• The limit of using purged CuO, CaO and Al2O3 for clinker production needs to
be revised, preferably with the support of experimental analyses.
• Possibility of separating CuO/Al2O3 sorbent from the purge stream would allow
significant reductions in the variable cost as presented in Chapter 4. Thus, it is
worth to investigate potential technologies for this purpose.
• Further experimental studies would verify the stability and performance of the
CLC sorbent to support the idea of reusing purge sorbent in the Ca-Cu looping
process.
• The use of syngas (CO+H2) from gasification of solid fuels, such as coal,
petroleum coke and biomass as an alternative fuel to methane in the calciner
would reduce the variable cost of this system as mentioned in Chapter 4. A
follow-up study may be conducted to investigate this option.
• An alternative method of providing steam into the stripper rather than having an
on-site CHP plant, i.e. availability of steam source in nearby locations, would be
an option to reduce the energy consumption associated with the amine process.
• The application of the advanced solvent systems, requiring less energy
consumption for carbon capture than an amine process and having improved
tolerance to SOx and NOx, deserves to be investigated further.
• Although a remarkable amount of dual-stage membrane process schemes have
been investigated in this study, the number of possible configuration can still be
expanded to improve the energy consumption and cost requirements.
• Further studies needs to be conducted to improve the accuracy of the membrane
models in use.
• A detailed combustor-carbonator model would allow more accurate prediction of
combustion and carbonation efficiencies in this reactor.

184
References

Abad, A., Adanez, J., Garcia-Libiano F., de Diego, L.F., Gayan, P. and Celaya, J.,
2007. Mapping of the Range of Operational Conditions for Cu-, Fe-, and Ni-based
Oxygen Carriers in Chemical-looping Combustion. Chem. Eng. Sci. 62, 533–549.
Abanades, J.C., 2002. The maximum capture efficiency of CO2 using a
carbonation/calcination cycle of CaO/CaCO3. Chemical Engineering Journal 90,
303-306.
Abanades, J.C., Alvarez, D., 2003. Conversion limits in the reaction of CO2 with
lime. Energy&Fuels 17,308–316.
Abanades, J. C., Anthony, E. J., Lu, D. Y., Salvador, C., Alvarez, D., 2004. Capture
of CO2 from combustion gases in a fluidized bed of CaO. AIChE Journal 50(7),
1614-1622.
Abanades, J.C., Anthony, E.J., Wang, J., Oakey, J.E., 2005. Fluidized bed
combustion systems integrating CO2 capture with CaO. Environ Sci Technol. 39,
2861–2866.
Abanades, J. C., Grasa, G., Alonso, M., Rodriguez, N., Anthony, E. J., Romeo, L.
M., 2007. Cost structure of a postcombustion CO2 capture system using CaO.
Environmental science & technology 41(15), 5523-5527.
Abanades, J. C., Murillo, R., 2009. Method of Capturing CO2 by Means of CaO and
the Exothermal Reduction of a Solid. European Patent Application EP2305366
A1.
Abanades, J.C., Murillo, R., Fernandez, J.R., Grasa, G., Martinez, I., 2010. New CO2
Capture Process for Hydrogen Production Combining Ca and Cu Chemical
Loops. Environ. Sci. Technol. 44, 6901–6904.
Abanades, J. C., Alonso, M., Rodríguez, N., 2011a. Biomass Combustion with in
Situ CO2 Capture with CaO. I. Process Description and Economics. Industrial &
Engineering Chemistry Research 50(11), 6972-6981.
Abanades, J. C., Alonso, M., Rodriguez, N., 2011b. Experimental validation of in-
situ CO2 capture with CaO during the low temperature combustion of biomass in
a fluidized bed reactor. International Journal of Greenhouse Gas Control 5(3),
512-520.
Adanez, J., Abad, A., Garcia-Labiano, F., Gayan, P., de Diego, L. F., 2012. Progress
in chemical-looping combustion and reforming technologies. Progress in Energy
and Combustion Science 38(2), 215-282.
Ahn, H., Luberti, M., Liu, Z., Brandani, S., 2013. Process configuration studies of
the amine capture process for coal-fired power plants. International Journal of
Greenhouse Gas Control 16, 29-40.
Albrecht, K. O., 2008a. Development and testing of a combined catalyst/sorbent
core-in-shell material for the production of high concentration hydrogen. PhD
Thesis, Iowa State University.

185
Albrecht, K. O., Wagenbach, K. S., Satrio, J. A., Shanks, B. H., Wheelock, T. D.,
2008b. Development of a CaO-based CO2 sorbent with improved cyclic stability.
Industrial & Engineering Chemistry Research 47(20), 7841-7848.
Alonso, M., Rodríguez, N., Grasa, G., Abanades, J. C., 2009. Modelling of a
fluidized bed carbonator reactor to capture CO2 from a combustion flue gas.
Chemical Engineering Science 64(5), 883-891.
Alonso, M., Rodríguez, N., González, B., Grasa, G., Murillo, R., Abanades, J.C.,
2010. Carbon dioxide capture from combustion flue gases with a calcium oxide
chemical loop. Experimental results and process development. International
Journal of Greenhouse Gas Control 4, 167-173.
Alonso, M., Rodríguez, N., González, B., Arias, B., Abanades, J. C., 2011a. Biomass
combustion with in Situ CO2 capture by CaO. II. Experimental results. Industrial
& Engineering Chemistry Research 50(11), 6982-6989.
Alonso, M., Rodriguez, N., Gonzalez, B., Arias, B., Abanades, J. C., 2011b. Capture
of CO2 during low temperature biomass combustion in a fluidized bed using CaO.
Process description, experimental results and economics. Energy Procedia 4, 795-
802.
Alonso, M., Diego, M.E., Abanades, J.C., Perez, C., Chamberlain, J., 2013. In Situ
CO2 Capture with CaO in a 300 Kw Fluidized Bed Biomass Combustor, in TCCS-
7, Trondheim.
Alptekin, G.O., Jayaraman, A., Dubovik, M., Brickner, L., 2008. Oxygen Sorbents
for Oxy-Fuel Combustion. Presented at 2008 AIChE Annual Meeting,
Philadelphia, USA.
Alsop, P.A., Chen, H., Tseng, H., 2007. The Cement Plant Operations Handbook,
Tradeship Publications Ltd, Surrey, UK.
Arias, B., Abanades, J. C., Anthony, E. J., 2011. Model for self-reactivation of highly
sintered CaO particles during CO2 capture looping cycles. Energy & Fuels 25(4),
1926-1930.
Arias, B., Diego, M.E., Abanades, J.C., Lorenzo, M., Diaz, L., Martinez, D., Alvarez,
J., Sanchez-Biezma, A., 2013. Demonstration of steady state CO2 capture in a 1.7
MWth calcium looping pilot. International Journal of Greenhouse Gas Control 18,
237-245.
Bhatty, J.I., Miller, F.M., Boahn, R.P., 2011. Innovations in Portland Cement
Manufacturing, Portland Cement Assn; 2nd edition.
Baker, R.W., 2002. Future directions of membrane gas separation technology.
Industrial & Engineering Chemistry Research 41(6), 1393-1411.
Baker, R.W., 2004. Membrane technology and applications. John Wiley & Sons, Ltd.
Bandyopadhyay, A., 2011. Amine versus ammonia absorption of CO2 as a measure
of reducing GHG emission: a critical analysis. Clean Technologies and
Environmental Policy 13(2), 269-294.
Baxter, L., 2005. Biomass-coal co-combustion: opportunity for affordable renewable
energy. Fuel 84(10), 1295-1302.

186
Berstad, D., Anantharaman, R., Jordal, K., 2012. Post-combustion CO2 capture from
a natural gas combined cycle by CaO/CaCO3 looping. International Journal of
Greenhouse Gas Control 11, 25-33.
Bhatia, S. K., Perlmutter, D. D., 1983. Effect of the product layer on the kinetics of
the CO2‐lime reaction. AIChE Journal 29(1), 79-86.
Bi, X.T., Liu, X., 2012. High density and high solid flux CFB risers for steam
gasification of solids fuels. Fuel Processing Technology 91 (8), 915-920.
Blamey, J., Anthony, E.J., Wang, J., Fennell, P.S., 2010. The calcium looping cycle
for large-scale CO2 capture, Prog. Energy Combust. Sci. 36, 260–279.
Blamey, J., Paterson, N. P., Dugwell, D. R., Stevenson, P., Fennell, P. S., 2011.
Reactivation of a CaO-based sorbent for CO2 capture from stationary sources.
Proceedings of the Combustion Institute 33(2), 2673-2681.
Bocciardo, D., Ferrari, M.C., Brandani, S., 2012. Customising UniSim Design:
membrane separation and its application to carbon capture, in EMEA HUG 2012,
Istanbul.
Bocciardo, D., Ferrari, M.C., Brandani, S., 2013. Modelling and multi-stage design
of membrane processes applied to carbon capture in coal-fired power plants.
Energy Procedia 37, 932-940.
Bocciardo, D., 2014. Optimisation and integration of membrane processes in coal-
fired power plants with carbon capture and storage. PhD Thesis, University of
Edinburgh.
Bogue, R.H., 1929. Calculation of the Compounds in Portland Cement. Industrial &
Engineering Chemistry Analytical Edition 1 (4), 192-197.
Boot-Handford, M. E., Abanades, J. C., Anthony, E. J., Blunt, M. J., Brandani, S.,
Mac Dowell, N., ... Fennell, P. S, 2014. Carbon capture and storage update.
Energy & Environmental Science 7(1), 130-189.
Bosoaga, A., Masek, O., Oakey, J.E., 2009. CO2 capture technologies for cement
industry. Energy Procedia 1(1), 133-140.
Böhringer, C., Rutherford, T., Tol, R., 2009. The EU 20/20/20 targets; an overview
of the EMF assessment. Energy Economics 31, 268-273.
Broda, M., Kierzkowska, A. M., Müller, C. R., 2012. Application of the Sol–Gel
Technique to Develop Synthetic Calcium‐Based Sorbents with Excellent Carbon
Dioxide Capture Characteristics. ChemSusChem 5(2), 411-418.
CCSGI “CCS Global Institute”, 2010. The Status of CCS Projects Interim Report,
Canberra, Australia.
Chang, M. H., Huang, C. M., Liu, W. H., Chen, W. C., Cheng, J. Y., Chen, W., ...,
Hsu, H. W., 2013. Design and Experimental Investigation of Calcium Looping
Process for 3‐kWth and 1.9‐MWth Facilities. Chemical Engineering &
Technology 36(9), 1525-1532.

187
Charitos, A., Rodriguez, N., Hawthorne, C., Alonso, M., Zieba, M., Arias, B.,
Kopanakis, G.,Scheffknecht, G., Abanades, J.C., 2011.Experimental validation of
the calcium looping CO2 capture process with two circulating fluidized bed
carbonator reactors. Ind. Eng. Chem. Res. 50, 9685–9695.
Chrissafis, K., Dagounaki, C., Paraskevopoulos, K. M., 2005. The effects of
procedural variables on the maximum capture efficiency of CO2 using a
carbonation/calcination cycle of carbonate rocks. Thermochimica acta 428(1),
193-198.
CO2CRC, 2014. CO2 capture/separation technologies. Co2crc.com.au. Retrieved
February 26, 2014, from
(http://www.co2crc.com.au/aboutccs/cap_absorption.html).
Coppola, A., Montagnaro, F., Salatino, P., Scala, F., 2012. Fluidized bed calcium
looping: The effect of SO2 on sorbent attrition and CO2 capture capacity.
Chemical Engineering Journal 207, 445-449.
Curran, G. P., Fink, C. E., Gorin, E., 1967. Carbon dioxide-acceptor gasification
process: studies of acceptor properties. Advanced Chemistry Services 69, 141-
165.
CW Group, 2012. Global Cement Volume Forecast Report.
Dai, J., Sokhansanj, S., Grace, J. R., Bi, X., Lim, C. J., Melin, S., 2008. Overview
and some issues related to co‐firing biomass and coal. The Canadian Journal of
Chemical Engineering 86(3), 367-386.
de Diego, L. F., Garcı́a-Labiano, F., Adánez, J., Gayán, P., Abad, A., Corbella, B.
M., Marı́a Palacios, J., 2004. Development of Cu-based oxygen carriers for
chemical-looping combustion. Fuel 83(13), 1749-1757.
Dean, C. C., Blamey, J., Florin, N. H., Al-Jeboori, M. J., Fennell, P. S., 2011a. The
calcium looping cycle for CO2 capture from power generation, cement
manufacture and hydrogen production. Chemical Engineering Research and
Design 89(6), 836-855.
Dean, C. C., Dugwell, D., Fennell, P. S., 2011b. Investigation into potential synergy
between power generation, cement manufacture and CO2 abatement using the
calcium looping cycle. Energy & Environmental Science 4(6), 2050-2053.
DECC, 2012a. UK Electricity Generation Costs Update.
DECC, 2012b. Government response to the consultation on proposals for the levels
of banded support under the Renewables Obligation for the period 2013-17 and
the Renewable Obligation Order 2012. Presented to Parliament by the Secretary
of State for Energy and Climate Change: London, U.K.
DECC, 2014. DECC 2050 Pathways Calculator.
(http://2050-calculator-tool.decc.gov.uk/)
Demirbas, A., 2003. Sustainable cofiring of biomass with coal. Energy Conversion
and Management 44(9), 1465-1479.

188
Demirbas, M. F., Balat, M., Balat, H., 2009. Potential contribution of biomass to the
sustainable energy development. Energy Conversion and Management 50(7),
1746-1760.
Dieter, H., Bidwe, A.R., Varela-Duelli, G., Charitos, A., Hawthorne, C.,
Scheffknecht, G., 2014. Development of the calcium looping CO2 capture
technology from lab to pilot scale at IFK, University of Stuttgart. Fuel (127), 23-
37.
Dlugokencky, E., Tans, P., 2014, NOAA/ESRL.
(http://www.esrl.noaa.gov/gmd/ccgg/trends/)
DOE, 1999. Carbon sequestration: research and development. U.S. Department of
Energy Report, Office of Science, Office of Fossil Energy, U.S. Department of
Energy.
DOE, 2003. Greenhouse Gas Emissions Control By Oxygen Firing In Circulating
Fluidized Bed Boilers: Phase 1 - a Preliminary Systems Evaluation. Alstom Power
Inc. and National Energy Technology Laboratory, US Department of Energy.
DOE, 2007. Cost and performance baseline for fossil energy plants. Volume 1:
Bituminous coal and natural gas to electricity final report. National Energy
Technology Laboratory.
DOE, 2010. Cost and performance for Low-Rank Pulverized Coal Oxycombustion
Energy Plants. Final Report, NETL-401/093010.
Dong, R.F., Lu, H.F., Yu, Y.S., Zhang, Z.X., 2012. A feasible process for
simultaneous removal of CO2, SO2 and NOx in the cement industry by NH3
scrubbing. Applied Energy 97, 185–191.
ECRA, 2007. Technical Report: Carbon Capture Technology – Options and
Potentials for the Cement Industry.
ECRA, 2009. Technical Report: ECRA CCS Project – Report about Phase II.
ECRA, 2012. Technical Report: ECRA CCS Project – Report on Phase III.
EG Science “EU Climate Change Expert Group”, 2008. The 2°C target Information
Reference Document.
(http://www.climateemergencyinstitute.com/uploads/2C_EU.pdf)
EIA, U.S. Energy Information Administration, 2013.
(http://www.eia.gov/dnav/ng/hist/n3035us3m.htm)
EPRI, 2009. Updated Cost and Performance Estimates for Clean Coal Technologies
Including CO2 Capture – 2009, Electric Power Research Institute, Palo Alto,
California.
Faaij, A., Meuleman, B., Turkenburg, W., van Wijk, A. D., Bauen, A., Rosillo-Calle,
F., Hall, D., 1998. Externalities of biomass based electricity production compared
with power generation from coal in the Netherlands. Biomass and Bioenergy
14(2), 125-147.
Favre, E., 2011. Membrane processes and postcombustion carbon dioxide capture:
challenges and prospects. Chemical Engineering Journal 171(3), 782-793.

189
Fennell, P. S., Davidson, J. F., Dennis, J. S., Hayhurst, A. N., 2007. Regeneration of
sintered limestone sorbents for the sequestration of CO2, from combustion and
other systems. J. Energy Inst. 80, 116–119.
Fernández, J. R., Abanades, J. C., Murillo, R., Grasa, G., 2012. Conceptual design of
a hydrogen production process from natural gas with CO2 capture using a Ca–Cu
chemical loop. International Journal of Greenhouse Gas Control 6, 126-141.
Figueroa, J. D., Fout, T., Plasynski, S., McIlvried, H., Srivastava, R. D., 2008.
Advances in CO2 capture technology—The US Department of Energy's Carbon
Sequestration Program. International Journal of Greenhouse Gas Control 2(1), 9-
20.
Finkenrath, M., 2011. Cost and performance of carbon dioxide capture from power
generation (No. 2011/5). OECD Publishing.
Fisher, K. S., Beitler, C., Rueter, C., Searcy, K., Rochelle, G., Jassim, M., 2005.
Integrating MEA Regeneration with CO2 Compression and Peaking to Reduce
CO2 Capture Costs (No. DOE/ER/84111). Trimeric Corporation.
Florin, N., Fennell, P., 2010. Carbon capture technology: future fossil fuel use and
mitigating climate change. Grantham Institute for Climate Change, briefing paper,
(3).
Forero, C. R., Gayán, P., García-Labiano, F., De Diego, L. F., Abad, A., Adánez, J.,
2011. High temperature behaviour of a CuO/γAl2O3 oxygen carrier for chemical-
looping combustion. International Journal of Greenhouse Gas Control 5(4), 659-
667.
Garcia-Labiano, F., Abad, A., de Diego, L.F., Gayan, P., Adanez, J., 2002.
Calcination of calcium-based sorbents at pressure in a broad range of CO2
concentrations. Chemical Engineering Science 57, 2381-2393.
García-Labiano, F., De Diego, L. F., Adánez, J., Abad, A., Gayán, P., 2004.
Reduction and oxidation kinetics of a copper-based oxygen carrier prepared by
impregnation for chemical-looping combustion. Industrial & engineering
chemistry research 43(26), 8168-8177.
Grasa, G.S., Abanades, J.C., 2006. CO2 capture capacity of CaO in long series of
carbonation/calcinations cycles. Industrial & Engineering Chemistry Research 45,
8846–8851.
Grasa, G., González, B., Alonso, M., Abanades, J. C., 2007. Comparison of CaO-
based synthetic CO2 sorbents under realistic calcination conditions. Energy &
Fuels 21(6), 3560-3562.
Grasa, G.S., Abanades, J.C., Alonso, M., Gonzalez, B., 2008a. Reactivity of highly
cycled particles of CaO in a carbonation/calcination loop. Chemical Engineering
Journal 137 (3), 561 – 567.
Grasa, G. S., Alonso, M., Abanades, J. C., 2008b. Sulfation of CaO particles in a
carbonation/calcination loop to capture CO2. Industrial & Engineering Chemistry
Research 47(5), 1630-1635.
Harlick, P.J.E., Tezel, F.H., 2004. An experimental adsorbent screening study for
CO2 removal from N2. Microporous and Mesoporous Materials 76, 71–76.

190
Harrison, D. P., 2008. Sorption-Enhanced Hydrogen Production: A Review. Ind.
Eng. Chem. Res. 47, 6486–6501.
Hasanbeigi, A., Price, L., Lin, E., 2012. Emerging energy-efficiency and CO2
emission-reduction technologies for cement and concrete production: a technical
review. Renewable and Sustainable Energy Reviews 16, 6220–6238.
Hassan, S.M.N., 2005. Techno-economic study of CO2 capture process for cement
plants. Masters Thesis, Chemical Engineering. University of Waterloo, Canada.
He, X., Hagg, M. B., 2012. Membranes for environmentally friendly energy
processes. Membranes 2(4), 706-726.
Hiltunen, M., Barišić, V., Coda Zabetta, E., 2008. Combustion of different types of
biomass in CFB boilers. In 16th European Biomass Conference.
Honeywell, 2010. Unisim Design Simulation Basis Reference Guide, R400 Release;
Honeywell: London.
Hossain, M. M., de Lasa, H. I., 2008. Chemical-Looping Combustion (CLC) for
Inherent CO2 Separations - A Review. Chem. Eng. Sci. 63, 4433–4451.
Hotta, A., 2010. Foster wheeler’s solutions for large scale CFB boiler technology:
features and operational performance of Łagisza 460 MWe CFB boiler. In
Proceedings of the 20th international conference on fluidized bed combustion (pp.
59-70). Springer Berlin Heidelberg.
Hu, X., Mangano, E., Friedrich, D., Ahn, H., Brandani, S., 2014. Diffusion
mechanism of CO2 in 13X zeolite beads. Adsorption 20(1), 121-135.
Huang, C. H., Chang, K. P., Yu, C. T., Chiang, P. C., Wang, C. F., 2010.
Development of high-temperature CO2 sorbents made of CaO-based mesoporous
silica. Chemical Engineering Journal 161(1), 129-135.
Hurst, T. F., Cockerill, T. T., Florin, N. H., 2012. Life cycle greenhouse gas
assessment of a coal-fired power station with calcium looping CO2 capture and
offshore geological storage. Energy & Environmental Science 5(5), 7132-7150.
IEA, 2007. Greenhouse Gas R&D Programme. Improved Oxygen production
technologies. IEA report 2007/14.
IEA, 2008. CO2 Capture in the Cement Industry, July 2008/3.
IEA, 2009a. World Energy Outlook 2009.
(http://www.worldenergyoutlook.org/publications/weo-2009/)
IEA, 2009b. Carbon Emission Reductions up to 2050. In Cement Technology
Roadmap 2009.
IEA, 2009c. Energy Technology Transitions for Industry: Strategies for the Next
Industrial Revolution. International Energy Agency: Paris.
IEA, 2010. Energy Technology Perspectives 2010: Scenarios & Strategies to 2050,
Paris, France.
IEA, 2012. World Energy Outlook 2012.
(http://www.worldenergyoutlook.org/publications/weo-2012/).

191
IEA Clean Coal Center, 2011. CO2 abatement in the cement industry.
IEA GHG, 2003. Technical & Financial Assessment Criteria. Revision B2.
IEA GHG, 2009. Biomass CCS Study: Techno-Economic Evaluation of Biomass
Fired or Co-Fired Power Plant with Post-Combustion CO2 Capture, IEAGHG:
Cheltenham, U.K.
IECM “Integrate Environmental Control Model”, 2012. Technical Documentation:
Membrane-based CO2 Capture Systems for Coal-fired Power Plants. Carnegie
Mellon University, Pittsburgh, USA.
Imtiaz, Q., Kierzkowska, A. M., Müller, C. R., 2012. Coprecipitated, Copper‐Based,
Alumina‐Stabilized Materials for Carbon Dioxide Capture by Chemical Looping
Combustion. ChemSusChem 5(8), 1610-1618.
INDMIN, Industrial Minerals, 2013. (www.indmin.com/)
IPCC, 2005. Carbon Dioxide Capture and Storage, Cambridge University Press, UK.
IPCC, 2007. Climate Change 2007: Mitigation. Contribution of Working Group III
to the Fourth Assessment Report of the Intergovernmental Panel on Climate
Change. Cambridge University Press, Cambridge, United Kingdom and New
York, NY, USA.
IPCC, 2014. Climate Change 2013: The Physical Science Basis. Contribution of
Working Group I to the Fifth Assessment Report of the Intergovernmental Panel
on Climate Change. Cambridge: Cambridge University Press, at press.
Ishida, M., Jin, H., 1994. A New Advanced Power-Generation System Using
Chemical-Looping Combustion. Energy 19 (4), 415–422.
Ishitani, H., Johansson, T.B., 1996. Energy supply mitigation options, in Climate
Change 1995: Impacts, Adaptations, and Mitigation of Climate Change:
Scientific-Technical Analyses; Cambridge University Press: Cambridge, UK.
Jäntti, T., Nuortimo, K., 2012. Foster Wheeler Advanced Bio CFB Technology for
Large Scale Biomass & Peat Firing Power Plants. Presented at Russia Power,
Moscow, Russia.
Jaramillo, P., Griffin, W. M., McCoy, S. T., 2009. Life cycle inventory of CO2 in an
enhanced oil recovery system. Environmental science & technology 43(21), 8027-
8032.
Jared, C.P., Sean, P.I., Lynn, B., Andrew, J., Gregson, V., Shiaoguo, C., 2010.
DOE/NETL advanced carbon dioxide capture R&D program: technology update,
September 2010.
Kapetaki, Z., Ahn, H., Brandani, S., 2013. Detailed process simulation of pre-
combustion IGCC plants using coal-slurry and dry coal gasifiers. Energy Procedia
37, 2196-2203.
Khan, A. A., De Jong, W., Jansens, P. J., Spliethoff, H., 2009. Biomass combustion
in fluidized bed boilers: Potential problems and remedies. Fuel processing
technology 90(1), 21-50.

192
Khinast, J., Krammer, G. F., Brunner, C., Staudinger, G., 1996. Decomposition of
limestone: The influence of CO2 and particle size on the reaction rate. Chemical
engineering science 51(4), 623-634.
Kierzkowska, A.M., Müller, C.R., 2012. Development of calcium-based, copper-
functionalised CO2 sorbents to integrate chemical looping combustion into
calcium looping, Energy Environ. Sci. 5, 6061-6065.
Kierzkowska, A.M., Müller, C.R., 2013. Sol–gel derived, calcium-based, copper-
functionalised CO2 sorbents for an integrated chemical looping combustion-
calcium looping CO2 capture process ChemPlusChem 78, 92–100.
Kolovos, K., Barafaka, S., Kakali, G., Tsivilis, S., 2005. CuO and ZnO addition in
the cement raw mix: effect on clinkering process and cement hydration and
properties. Ceramics 49, 205–212.
Koppatz, S., Pfeifer, C., Rauch, R., Hofbauer, H., Marquard-Moellenstedt, T.,
Specht, M., 2009. H2 rich product gas by steam gasification of biomass with in
situ CO2 absorption in a dual fluidized bed system of 8 MW fuel input. Fuel
Process. Technol. 90, 914-921.
Kunii, D., Levenspiel, O., 1991. Fluidization Engineering, second ed. Butterworth-
Heinemann, Boston.
Kuramochi, T., Ramírez, A., Turkenburg, W., Faaij, A., 2012. Comparative
assessment of CO2 capture technologies for carbon-intensive industrial processes.
Progress in energy and combustion science 38(1), 87-112.
Kyoto, 1997. Kyoto Protocol to the United Nations Framework Convention on
Climate Change, Kyoto, 11 December 1997.
Lara, Y., Lisbona, P., Martínez, A., Romeo, L. M., 2013. Design and analysis of heat
exchanger networks for integrated Ca-looping systems. Applied Energy 111, 690-
700.
Lasheras, A., Ströhle, J., Galloy, A., Epple, B., 2011. Carbonate looping process
simulation using a 1D fluidized bed model for the carbonator. International
Journal of Greenhouse Gas Control 5(4), 686-693.
Leckner, B., 1998. Fluidized bed combustion: mixing and pollutant limitation.
Progress in Energy and Combustion Science 24(1), 31-61.
Leion, H., Mattisson, T., Lyngfelt, A., 2009. Use of ores and industrial products as
oxygen carriers in chemical-looping combustion. Energy & Fuels 23(4), 2307-
2315.
Li, Z-s., Cai, N-s, Croiset, E., 2008. Process analysis of CO2 capture from flue gas
using carbonation/calcination cycles. AlChE J 54, 1912-1925.
Li, J., Tharakan, P., Macdonald, D., Liang, X., 2013. Technological, economic and
financial prospects of carbon dioxide capture in the cement industry. Energy
Policy 61, 1377-1387.
Liang, X., Li, J., 2012. Assessing the value of retrofitting cement plants for carbon
capture: A case study of a cement plant in Guangdong, China. Energy Conversion
and Management 64, 454-465.

193
Linnhoff, B., Flower, J. R., 1978. Synthesis of heat exchanger networks: I.
Systematic generation of energy optimal networks. AIChE Journal 24(4), 633-
642.
LME, London Metal Exchange, 2013. (http://www.lme.com/)
Lu, H., Reddy, E.P., Smirniotis, P.G., 2006. Calcium oxide based sorbents for
capture of carbon dioxide at high temperatures. Ind. Eng. Chem. Res. 45, 3944–
3949.
Lu, D. Y., Hughes, R. W., Anthony, E. J., 2008. Ca-based sorbent looping
combustion for CO2 capture in pilot-scale dual fluidized beds. Fuel Processing
Technology 89(12), 1386-1395.
Lyon, R. K., Cole, J. A., 2000. Unmixed Combustion: An Alternative to Fire.
Combust. Flame 121, 249–261.
Lysikov, A. I., Salanov, A. N., Okunev, A. G., 2007. Change of CO2 carrying
capacity of CaO in isothermal recarbonation-decomposition cycles. Industrial &
Engineering Chemistry Research 46(13), 4633-4638.
Ma, X., Chen, H., Wang, P., 2010. Effect of CuO on the formation of clinker
minerals and the hydration properties. Cement and Concrete Research 40, 1681–
1687.
Maciejewska, A., Veringa, H., Sanders, J., Peteves, S. D., 2006. Co-firing of biomass
with coal: constraints and role of biomass pre-treatment. Petten, The Netherlands:
Institute for Energy 113, 100.
MacKenzie, A., Granatstein, D. L., Anthony, E. J., Abanades, J. C., 2007. Economics
of CO2 capture using the calcium cycle with a pressurized fluidized bed
combustor. Energy & fuels 21(2), 920-926.
Manovic, V., Anthony, E. J., 2007. Steam reactivation of spent CaO-based sorbent
for multiple CO2 capture cycles. Environmental science & technology 41(4),
1420-1425.
Manovic, V., Anthony, E. J., 2008. Thermal activation of CaO-based sorbent and
self-reactivation during CO2 capture looping cycles. Environmental science &
technology 42(11), 4170-4174.
Manovic, V., Lu, D., Anthony, E. J., 2008. Steam hydration of sorbents from a dual
fluidized bed CO2 looping cycle reactor. Fuel 87(15), 3344-3352.
Manovic, V., Anthony, E. J., 2009a. Screening of Binders for Pelletization of CaO-
Based Sorbents for CO2 Capture. Energy & Fuels 23(10), 4797-4804.
Manovic, V., Anthony, E.J., 2009b. CaO-based pellets supported by calcium
aluminate cements for high-temperature CO2 capture. Environ. Sci. Technol. 43,
7117-7122.
Manovic, V., Anthony, E. J., 2009c. Long-term behavior of CaO-based pellets
supported by calcium aluminate cements in a long series of CO2 capture cycles.
Industrial & Engineering Chemistry Research 48(19), 8906-8912.

194
Manovic, V., Anthony, E. J., 2010. CO2 carrying behavior of calcium aluminate
pellets under high-temperature/high-CO2 concentration calcination conditions.
Industrial & Engineering Chemistry Research 49(15), 6916-6922.
Manovic, V., Anthony, E. J., 2011a. Reactivation and remaking of calcium aluminate
pellets for CO2 capture. Fuel 90(1), 233-239.
Manovic, V., Anthony, E. J., 2011b. CaO-Based Pellets with Oxygen Carriers and
Catalysts. Energy Fuels 25 (10), 4846 – 4853.
Manovic, V., Wu, Y., He, I., Anthony, E. J., 2011c. Core-in-shell CaO/CuO-based
composite for CO2 capture. Ind. Eng. Chem. Res. 50, 12384−12391.
Manovic, V., Anthony, E. J., 2011d. Integration of Calcium and Chemical Looping
Combustion using Composite CaO/CuO-Based Materials. Environ. Sci. Technol.
45, 10750−10756.
Martinez, I., Grasa, G., Murillo, R., Arias, B., Abanades, J.C., 2010. Kinetics of
Calcination of Partially Carbonated Particles in a Ca-looping System for CO2
Capture. Energy&Fuels 26 (2), 1432-1440.
Martinez, I., Romano, M. C., Fernández, J. R., Chiesa, P., Murillo, R., Abanades, J.
C., 2014. Process design of a hydrogen production plant from natural gas with
CO2 capture based on a novel Ca/Cu chemical loop. Applied Energy 114, 192-
208.
McIlveen-Wright, D. R., Huang, Y., Rezvani, S., Mondol, J. D., Redpath, D.,
Anderson, M., Williams, B. C., 2011. A techno-economic assessment of the
reduction of carbon dioxide emissions through the use of biomass co-combustion.
Fuel 90(1), 11-18.
Mellows-Facer, A., 2010. Key issues for the new parliament. House of Commons
Library Research. (http://goo.gl/Gdv1n0)
Merkel, T. C., Lin, H., Wei, X., Baker, R., 2010. Power plant post-combustion
carbon dioxide capture: an opportunity for membranes. Journal of Membrane
Science 359(1), 126-139.
Merkel, T. C., Zhou, M., Baker, R. W., 2012. Carbon dioxide capture with
membranes at an IGCC power plant. Journal of Membrane Science 389, 441-450.
McCauley, K. J., Farzan, H., Alexander, K. C., McDonald, D. K., 2009.
Commercialization of Oxy-Coal Combustion: Applying Results of a Large 30
MWth Pilot Project. Energy Procedia 1, 439–446.
Naranjo, M., Brownlow, D. T., Garza, A., 2011. CO2 capture and sequestration in the
cement industry. Energy Procedia 4, 2716-2723.
NETL, 2012. Current and Future Technologies for Power Generation with Post-
Combustion Carbon Capture - Final Report, in DOE/NETL-2012/1557.
Nuortimo, K., 2013. Large scale CHP with CFB technology. Presented at 4th
European Conference of Renewable Heating and Cooling, Dublin, Ireland.
Obersteiner, M., Azar, C., Kauppi, P., Möllersten, K., Moreira, J., Nilsson, S., Van
Ypersele, J. P., 2001. Managing climate risk. Science 294(5543), 786-787.

195
Olivier, J. G., Janssens-Maenhout, G., Muntean, M., Peters, J. A., 2013. Trends in
global CO2 emissions 2013 report. PBL Netherlands Environmental Assessment
Agency.
Ozcan, D. C., 2010. Development of a sorbent for carbon dioxide. MSc Thesis, Iowa
State University.
Ozcan, D. C., Shanks, B. H., Wheelock, T. D., 2011. Improving the stability of a
CaO-based sorbent for CO2 by thermal pretreatment. Industrial & Engineering
Chemistry Research 50(11), 6933-6942.
Ozcan, D.C., Ahn, H., Kierzkowska, A.M., Müller, C.R., Brandani, S., 2013. Process
integration of chemical looping combustion into calcium looping for CO2 capture
from a cement plant, 5th HTSLCN Meeting, Cambridge, UK.
Pacciani, R., Müller, C. R., Davidson, J. F., Dennis, J. S., Hayhurst, A. N., 2008.
Synthetic Ca‐based solid sorbents suitable for capturing CO2 in a fluidized bed.
The Canadian journal of chemical engineering 86(3), 356-366.
Pan, X., Clodic, D., Toubassy, J., 2013. CO2 capture by antisublimation process and
its technical economic analysis. Greenhouse Gases: Science and Technology 3
(1), 8–20.
Pennline, H. W., Luebke, D. R., Jones, K. L., Myers, C. R., Morsi, B. I., Heintz, Y.
J., Ilconich, J. B., 2008. Progress in Carbon Dioxide Capture and Separation
Research for Gasification-Based Power Generation Point Sources. Fuel Process.
Technol. 89, 897–907.
Plötz, S., Bayrak, A., Galloy, A., Kremer, J., Orth, M., Wieczorek, M., Ströhle, J.,
Epple, B., 2012. First carbonate looping experiments with a 1 MWth test facility
consisting of two interconnected CFBs. In: 21st International conference on
fluidized bed combustion, Naples (Italy), 421–428.
Qin, C., Yin, J., Liu, W., An, H., Feng, B., 2012. Behavior of CaO/CuO based
composite in a combined calcium and copper chemical looping process. Ind. Eng.
Chem. Res. 51, 12274−12281.
Qin, C., Yin, J., Luo, C., An, H., Liu, W., Feng, B., 2013. Enhancing the
performance of CaO/CuO based composite for CO2 capture in a combined Ca-Cu
chemical looping process. Chemical Engineering Journal 228, 75-86.
Ramasubramanian, K., Verweij, H., Winston Ho, W. S., 2012. Membrane processes
for carbon capture from coal-fired power plant flue gas: A modeling and cost
study. Journal of Membrane Science 421, 299-310.
Rampinelli, G., 2010. Modello matematico di un reattore per la cattura della CO2
post-combustione tramite ossido di calico. Masters Thesis. Politecnico di Milano,
Italy.
Rao, A. B., Rubin, E. S. A., 2002. Technical, Economic, and Environmental
Assessment of Amine-Based CO2 Capture Technology for Power Plant
Greenhouse Gas Control. Environ. Sci. Technol. 36, 4467– 4475.
Reh, L., 1995. New and efficient high-temperature processes with circulating fluid
bed reactor. Chemical Engineering & Technology 18 (2), 75-89.

196
Rhodes, J. S., Keith, D. W., 2008. Biomass with capture: negative emissions within
social and environmental constraints: an editorial comment. Climatic Change
87(3), 321-328.
Robeson, L. M., 2008. The upper bound revisited. Journal of Membrane Science
320(1), 390-400.
Rodriguez, N., Alonso, M., Grasa, G., Abanades, J. C., 2008a. Heat requirements in a
calciner of CaCO3 integrated in a CO2 capture system using CaO. Chemical
Engineering Journal 138(1), 148-154.
Rodriguez, N., Alonso, M., Grasa, G., Abanades, J. C., 2008b. Process for capturing
CO2 arising from the calcination of the CaCO3 used in cement manufacture.
Environmental science & technology 42(18), 6980-6984.
Rodriguez, N., Alonso, M., Abanades, J.C., Grasa, G., Murillo, R., 2009. Analysis of
a process to capture the CO2 resulting for pre-calcination of the limestone feed to
a cement plant. Energy Procedia 1, 141–148.
Rodríguez, N., Alonso, M., Abanades, J. C., 2010. Average activity of CaO particles
in a calcium looping system. Chemical Engineering Journal 156(2), 388-394.
Rodríguez, N., Alonso, M., Abanades, J.C., 2011a. Experimental investigation of a
circulating fluidized-bed reactor to capture CO2 with CaO. AIChE J. 57, 1356-
1366.
Rodríguez, N., Murillo, R., Alonso, M., Martinez, I., Grasa, G., Abanades, J.C.,
2011b. Analysis of a process for capturing the CO2 resulting for pre-calcination of
limestone in a cement plant. Industrial & Engineering Chemistry Research 50 (4),
2126–2132.
Rodriguez, N., Murillo, R., Abanades, J.C., 2012. CO2 Capture from Cement Plants
Using Oxyfired Precalcination and/or Calcium Looping. Environmental Science
& Technology 46 (4), 2460-2466.
Romano, M. C., 2012. Modeling the carbonator of a Ca-looping process for CO2
capture from power plant flue gas. Chemical Engineering Science 69(1), 257-269.
Romano, M. C., Spinelli, M., Campanari, S., Consonni, S., Cinti, G., Marchi, M.,
Borgarello, E., 2013. The Calcium Looping Process for Low CO2 Emission
Cement and Power. Energy Procedia 37, 7091-7099.
Romeo, L.M., Abanades, J.C., Escosa, J.M., Pano, J., Gimenez A., 2008. Oxyfuel
carbonation/calcination cycle for low cost CO2 capture in existing power plants.
Energy Convers Manage 49:2809–2814.
Romeo, L.M., Lara, Y., Lisbona, P., Escosa, J.M., 2009. Optimizing make-up flow in
a CO2 capture system using CaO. Chem Eng J 147, 252–258.
Romeo, L.M., Catalina, D., Lisbona, P., Lara, Y., Martínez, A., 2011. Reduction of
greenhouse gas emissions by integration of cement plants, power plants, and CO2
capture systems. Greenhouse Gases: Science and Technology 1, 72–82.
Rong, N., Wang, Q., Fang, M., Cheng, L., Luo, Z., Cen, K., 2013. Steam Hydration
Reactivation of CaO-Based Sorbent in Cyclic Carbonation/Calcination for CO2
Capture. Energy & Fuels 27(9), 5332-5340.

197
Rubin, E. S., Yeh, S., Antes, M., Berkenpas, M., Davison, J., 2007. Use of
experience curves to estimate the future cost of power plants with CO2 capture.
International journal of greenhouse gas control 1(2), 188-197.
Ryu, H. J., Grace, J. R., Lim, C. J., 2006. Simultaneous CO2/SO2 capture
characteristics of three limestones in a fluidized-bed reactor. Energy & fuels
20(4), 1621-1628.
Sakadjian, B. B., Iyer, M. V., Gupta, H., Fan, L. S., 2007. Kinetics and structural
characterization of calcium-based sorbents calcined under subatmospheric
conditions for the high-temperature CO2 capture process. Industrial &
Engineering Chemistry Research 46(1), 35-42.
Salvador, C., Lu, D., Anthony, E. J., Abanades, J. C., 2003. Enhancement of CaO for
CO2 capture in an FBC environment. Chemical Engineering Journal 96(1), 187-
195.
Schaeffer, G. J., Boots, M. G., Mitchell, C., Timpe, C., Cames, M., Anderson, T.,
2013. The implications of tradable green certificates for the deployment of
renewable electricity: mid-term report. Policy Studies 2012, 2011.
Scholes, C. A., Chen, G. Q., Tao, W. X., Bacus, J., Anderson, C., Stevens, G. W.,
Kentish, S. E., 2011. The effects of minor components on the gas separation
performance of membranes for carbon capture. Energy Procedia 4, 681-687.
SETIS, 2013. Calculation of Energy Production Costs. Institute for Energy,
Directorate-General Joint Research Centre: Petten, The Netherlands.
(http://setis.ec.europa.eu/EnergyCalculator/)
Shimizu, T., Hirama, T., Hosoda, H., Kitano, K., Inagaki, M., Tejima, K., 1999. A
twin fluid-bed reactor for removal of CO2 from combustion processes. Chem Eng
Res Des 77, 62–68.
Squires, A. M., 1967. Cyclic use of calcined dolomite to desulfurize fuels undergoing
gasification. Adv. Chem. Ser. 69, 205–229.
Silaban, A., Harrison, D.P., 1995. High-temperature capture of carbon dioxide:
characteristics of the reversible reaction between CaO (s) and CO2 (g). Chemical
Engineering Communications 137,177–190.
Silaban, A., Narcida, M., Harrison, D. P., 1996. Characteristics of the reversible
reaction between CO2(g) and calcined dolomite. Chemical Engineering
Communications 146: 149-162.
Soukup, G., Pfeifer, C., Kreuzeder, A., Hofbauer, H., 2009. In situ CO2 capture in a
dual fluidized bed biomass steam gasifier - Bed material and fuel variation.
Chemical Engineering and Technology 32, 348-354.
Srivastava, R. K., Jozewicz, W., Singer, C., 2001. SO2 scrubbing technologies: a
review. Environmental Progress 20(4), 219-228.
Stadler, H., Beggel, F., Habermehl, M., Persigehl, B., Kneer, R., Modigell, M.,
Jeschke, P., 2011. Oxyfuel coal combustion by efficient integration of oxygen
transport membranes. International Journal of Greenhouse Gas Control 5(1), 7-15.

198
Stallmann, O., 2013. Integrated carbon dioxide capture for cement plants. Publication
No. WO/2013/024340.
Stevens, D. J., 2001. Hot gas conditioning: recent progress with larger-scale biomass
gasification systems. NREL Subcontractor Report (NREL/SR-510-29952).
Sun, P., Grace, J. R., Lim, C. J., Anthony, E. J., 2007. Removal of CO2 by calcium-
based sorbents in the presence of SO2. Energy & fuels 21(1), 163-170.
Sun, P., Grace, J. R., Lim, C. J., Anthony, E. J., 2008. Investigation of Attempts to
Improve Cyclic CO2 Capture by Sorbent Hydration and Modification. Ind. Eng.
Chem. Res. 47, 2024–2032.
Sun, R., Li, Y., Liu, H., Wu, S., Lu, C., 2012. CO2 capture performance of calcium-
based sorbent doped with manganese salts during calcium looping cycle. Applied
Energy 89(1), 368-373.
Taylor, H.F.W., 1990. Cement Chemistry, Academic Press Ltd, New York.
Telesca, A., Calabrese, D., Marroccoli, M., Tomasulo, M., Valenti, G.L.,
Montagnaro, F., 2014. Spent limestone sorbent from calcium looping cycle as a
raw material for the cement industry. Fuel 118, 202-205.
Torp, T., Gale, J., 2004. Demonstrating storage of CO2 in geological reservoirs: The
Sleipner and SACS projects. Energy 29, 1361– 1369.
Tuinier, M. J., Hamers, H. P., van Sint Annaland, M., 2011. Techno-economic
evaluation of cryogenic CO2 capture - A comparison with absorption and
membrane technology. International Journal of Greenhouse Gas Control 5(6),
1559-1565.
Turton, R., Bailie, R.C., Whiting, W.B., Shaeiwitz, J.A., Bhattacharyya, D., 2009.
Analysis, Synthesis, and Design of Chemical Processes (third ed.), Prentice Hall.
U. K. Parliament (UKP), 2008. Climate change act 2008. London, UK.
United Nations Framework Convention on Climate Change (UNFCC), 2007. Article
2, Bali 3–15 December 2007.
Vera, E.R.M., 2009. Method for capturing CO2 produced by cement plants by using
the calcium cycle. Publication No. US20090255444 A1.
Wang, Y., Lin, S., Suzuki, Y., 2008. Limestone calcination with CO2 capture (II):
Decomposition in CO2/steam and CO2/N2 atmospheres. Energy & Fuels 22(4),
2326-2331.
Wang, W., Ramkumar, S., Wong, D., Fan, L. S., 2012. Simulations and process
analysis of the carbonation–calcination reaction process with intermediate
hydration. Fuel 92(1), 94-106.
WBCSD, 2009. Cement Industry and CO2 Performance. ‘Getting the Numbers
Right’, Cement Sustainability Initiative, Geneva.
Weimer, T., Berger, R., Hawthorne, C., Abanades, J. C., 2008. Lime enhanced
gasification of solid fuels: Examination of a process for simultaneous hydrogen
production and CO2 capture. Fuel 87(8), 1678-1686.

199
Werther, J., Saenger, M., Hartge, E. U., Ogada, T., Siagi, Z., 2000. Combustion of
agricultural residues. Progress in energy and combustion science 26(1), 1-27.
WWF, 2008. Blueprint for the Cement Industry: How to Turn Around the Trend of
Cement Related Emissions in the Developing World.
Xu, G., Li, L., Yang, Y., Tian, L., Liu, T., Zhang, K., 2012. A novel CO2 cryogenic
liquefaction and separation system. Energy 42(1), 522-529.
ZEP, 2011. The Cost of CO2 Capture, Transport and Storage. Post-demonstration
CCS in the EU, Technology Platform for Zero Emission Fossil Fuel Power Plants,
Brussels.
Zhai, H., Rubin, E. S., 2013. Techno-Economic Assessment of Polymer Membrane
Systems for Postcombustion Carbon Capture at Coal-Fired Power Plants.
Environmental science & technology 47(6), 3006-3014.
Zhao, L., Riensche, E., Blum, L., Stolten, D., 2010. Multi-stage gas separation
membrane processes used in post-combustion capture: Energetic and economic
analyses. Journal of Membrane Science 359(1), 160-172.

200
Appendix A

Mass and energy balance calculations for the base cement

plant, Ca-looping process and preliminary steam cycle

design detailed in Chapter 2

201
31

To
Atmosphere

To
Collected
Atmosphere B/F
Gas Flow Dust
3
Solid Flow 27
Coal To
B/F Pet Coke F/D 26 B/F
28
Primary Atmosphere
9 Air
29 30 25
22
Secondary Excess
5 8 11 13 15 19 Air
Raw 1 1st 2nd 3rd 4th Air
R/M Pre-C Kiln Cooler
Meal PHE PHE PHE PHE Clinker
4 7 10 12 14 18 21
24
2 6 17 20 23
Air Air Air Air Cooling
Air
16 Tertiary Air
Figure A1 Schematic diagram of a cement plant without a CO2 capture unit (Base Case) (IEA, 2008). Abbreviations: R/M, Raw Mill; B/F, Bag Filter;
F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner.
202
Table A1 Stream compositions for base cement plant (only two digits after decimal have been shown).
Stream 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
Number
Stream Name Raw Air B/F R/M Preheater Air 1st 2nd PHE 1st PHE Exit 2nd 3rd PHE 3rd 4th PHE 4th Pre-C Tertiary Air
meal In- Exit Exit Exit Gas In- PHE Exit Gas to Fuel PHE Exit PHE Exit PHE Exit Air In-
leak Gas leak Exit Gas Drying Exit Gas Exit Gas Exit Gas leak
Component
CO2 kg/s 23.97 23.97 25.73 1.76 25.73 25.73 23.66
CaCO3 kg/s 37.94 40.49 2.55 42.86 5.10 0.19 45.91 8.49 46.24 8.49 33.94 0.90
CaSO4 kg/s 0.01 0.01 0.08 0.08 0.33 0.33
H2 kg/s
O2 kg/s 3.75 5.06 1.31 0.97 0.49 0.1 0.49 0.49 0.49 3.81 1.13
SO2 kg/s 0.1 0.11
N2 kg/s 12.2 40.45 28.23 3.18 27.13 2.07 27.13 27.13 27.16 12.42 3.68
Ash kg/s 0.02 0.03 0.03 0.11 0.11
C kg/s
S kg/s 0.16 0.17 0.12 0.01 0.13 0.02 0.03 0.03 0.11
H2O kg/s 4.19 5.87 1.67 1.80 0.12 1.79 1.79 1.62
C2 S kg/s 0.07 0.07 0.67 0.67 3.79 3.79 15.16 6.48
C3 A kg/s 0.05 0.05 0.19 0.19
C3 S kg/s 0.05 0.05 0.26 0.26 1.04 1.04
C4AF 0.04 0.04 0.14 0.14
Fe2O3 kg/s 0.03 0.03 0.15 0.15 0.60 0.31
Al2O3 kg/s 0.06 0.06 0.33 0.33 1.30 0.66
SiO2 kg/s 3.16 3.37 0.21 3.57 0.42 0.02 3.78 0.63 3.59 0.44 1.76 0.34
AS4H kg/s 4.74 5.06 0.32 5.35 0.63 0.02 5.69 0.97 5.49 0.77 3.07
AS2H2 kg/s 1.27 1.35 0.08 1.43 0.17 1.52 0.26 1.46 0.21 0.82
2FeO⋅OH kg/s 0.95 1.01 0.06 1.07 0.13 1.14 0.19 1.10 0.15 0.62
CaO kg/s 3.01
Total Flow kg/s 52.41 15.96 75.44 51.46 58.52 4.15 54.48 61.7 4.3 58.99 66.21 62.72 69.94 59.20 66.46 16.23 4.81
Temperature °C 9 9 110 110 309 9 309 490 309 490 634 634 760 760 915 908 9
Pressure atm - 1 1 - 1 1 - 1 1 - 1 - 1 - 1 1 1
Total Stream GJ/h -2415 -0.95 -1044 -2310 -938.5 -0.25 -2405 -1101 -68.96 -2561 -1256 -2689 -1382 -2524 -1217 +56.64 -0.29
Enthalpy
203
Stream 18-1 18-2 19-1 19-2 20 21-1 21-2 22-1 22-2 23 24 25 26 27 28 29 30 31
Number
Stream Name Pre-C Pre-C Kiln Kiln Air Kiln Kiln Secondary Secondary Cooling Clinker B/F Primary Coal Pet- Dry Dry B/F
Exit Exit Exit Exit In- Exit Exit Air Air Air Exit Air coke Coal Pet- Exit
Gas Gas leak Gas Coke Gas
Component
CO2 kg/s 4.82 4.82 1.76
CaCO3 kg/s 2.51 2.51 0.19
CaSO4 kg/s 0.92 0.92 0.12 0.12 0.80 0.80 0.80
H2 kg/s 0.10 0.04 0.10 0.04 2.07
O2 kg/s 0.31 0.31 0.38 2.65 2.65 16.78 10.33 0.36 0.16 0.02 0.16 0.02 0.01
SO2 kg/s 0.12 0.12
N2 kg/s 11.03 11.03 1.22 8.64 8.64 54.73 33.67 1.14 0.03 0.02 0.03 0.02
Ash kg/s 0.32 0.32 0.04 0.04 0.28 0.28 0.28 0.28 0.28
C kg/s 1.46 1.01 1.46 1.01
S kg/s 0.02 0.06 0.02 0.06
H2O kg/s 0.37 0.37 0.21 0.02 0.36
C2 S kg/s 17.94 17.94 0.82 0.82 5.42 5.43 5.43
C3 A kg/s 0.52 0.52 0.52 0.52 3.44 3.44 3.44
C3 S kg/s 2.89 2.89 2.89 2.89 19.08 19.08 19.08
C4AF 0.39 0.39 0.39 0.39 2.59 2.59 2.59
Fe2O3 kg/s 0.85 0.85
Al2O3 kg/s 1.84 1.84
SiO2 kg/s 0.93 0.93 0.02
AS4H kg/s 0.02
AS2H2 kg/s
2FeO⋅OH kg/s
CaO kg/s 8.36 8.36
Total Flow kg/s 37.5 37.5 21.44 21.44 1.60 31.61 31.61 11.29 11.29 71.52 31.61 44.00 1.50 2.26 1.18 2.29 1.16 4.30
Temperature °C 915 1250 1450 1025 9 1450 1370 1025 1245 9 60 279 9 9 9 110 110 110
Pressure atm - - 1 1 1 - - 1 1 1 - 1 1 - - - - 1
Total Stream MJ/h -1582 -1541 -264.3 -305.5 -0.09 -1260 -1270 +45.0 +55.71 -4.2 -1424 41.73 -0.09 -25.6 +0.52 - +2.21 -73.50
Enthalpy 22.73
204
Excess 42
Air

Gas Flow Steam CO2 Compressed


ASU Cycle Comp. CO2
Solid Flow 43
Oxygen 38
Heat stream
Make-up To
Qcarbonator 39 31
CaCO3 Atmosphere
37
Pet 40
Coke
B/F Collected
To Dust
36 32
Atmosphere
CaCO3 28
Pet Coke
3 Carb Cal Coal F/D 30
CaO 27
26
B/F 35 Purge 33 29 Primary B/F
9 34 11 Air
22
Secondary Excess 25
5 8 13 15 Air Air
Raw 1 1st 2nd 3rd 4th 19
R/M Pre-C Kiln Cooler
Meal PHE PHE PHE PHE
4 7 10 12 14 18 21 Clinker
24
2 6 17 20 23
Air Air Air Air Cooling
Air
16 Tertiary Air

Figure A2 Schematic diagram of the proposed process integration of a cement plant with a Ca-looping unit. Abbreviations: R/M, Raw Mill; B/F, Bag
Filter; F/D, Fuel Drying; PHE, Preheater; Pre-C, Pre-calciner; Carb, Carbonator; Calc, Calciner; ASU, Air Separation Unit; CO2 Comp, CO2
Compression.
205
Table A2 Stream compositions for cement plant with a Ca-looping unit at 1.65 F0/FCO2 (only two digits after decimal have been shown).
Stream Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
st nd st nd rd rd th th
Stream Name Raw Air B/F R/M Preheater Air 1 PHE 2 PHE 1 PHE Exit 2 PHE 3 PHE 3 PHE 4 PHE 4 PHE Pre-C
meal In- Exit Exit Exit Gas In- Exit Exit Gas Gas to Fuel Exit Exit Gas Exit Exit Gas Exit Exit Gas
leak Gas leak Drying
Component
CO2 kg/s 0.69 0.69 0.77 0.09 7.77 7.77 7.30
CaCO3 kg/s 8.62 9.17 0.55 9.70 1.15 0.07 10.38 1.85 10.46 1.92 7.68 0.21
CaSO4 kg/s 0.01 0.01 0.08 0.08 0.30 0.30
H2 kg/s
O2 kg/s 1.47 8.29 0.43 0.38 0.16 0.06 0.16 0.16 0.16
SO2 kg/s 0.1 0.10 0.01
N2 kg/s 4.79 37.38 11.76 1.25 12.02 1.50 12.02 12.02 12.02
Ash kg/s 0.04 0.04
C kg/s
S kg/s 0.16 0.17 0.12 0.01 0.13 0.02 0.13 0.03 0.10
H2O kg/s 1.64 2.55 0.91 1.02 0.12 1.02 1.02 0.85
C2 S kg/s 0.02 0.02 0.16 0.16 0.89 0.89 3.56 2.64
C3 A kg/s 0.05 0.05 0.19 0.19
C3 S kg/s 0.05 0.05 0.26 0.26 1.03 1.03
C4AF kg/s 0.04 0.04 0.14 0.14
Fe2O3 kg/s 0.03 0.03 0.15 0.15 0.60 0.31
Al2O3 kg/s 0.06 0.06 0.33 0.33 1.30 0.66
SiO2 kg/s 3.16 3.36 0.20 3.57 0.44 0.03 3.94 0.81 4.58 1.45 5.79 1.68
AS4H kg/s 4.74 5.04 0.30 5.33 0.63 0.04 5.66 0.96 5.46 0.77 3.06
AS2H2 kg/s 1.27 1.35 0.08 1.42 0.17 0.01 1.51 0.26 1.46 0.20 0.82
2FeO⋅OH kg/s 0.95 1.00 0.06 1.07 0.13 1.13 0.19 1.09 0.15 0.61
CaO kg/s
Total Flow kg/s 20.54 6.25 49.00 20.10 15.09 1.63 21.25 16.54 1.93 23.07 25.38 24.97 27.28 25.21 27.51
Temperature °C 9 9 110 110 291.4 9 291.4 456.2 291.4 456.2 588 588 701 701 915
Pressure atm - 1 >1 - >1 1 - >1 >1 - >1 - >1 - >1
Total Stream GJ/h -993 -0.37 -130.9 -949.6 -109.1 -0.1 -995 -167.9 -13.92 -1065 -453 --1137 -524.3 -1129 -516
Enthalpy
206
Stream 16 17 18-1 18-2 19-1 19-2 20 21-1 21-2 22-1 22-2 23 24 25 26 27
Number
Stream Name Tertiary Air Pre-C Pre-C Kiln Kiln Air Kiln Kiln Secondary Secondary Cooling Clinker Excess Primary Coal
Air In- Exit + Exit + Exit Exit In- Exit Exit Air Air Air Air to Air
leak Purge Purge Gas Gas leak R/M
Component
CO2 kg/s 2.63 2.63
CaCO3 kg/s 0.57 0.57
CaSO4 kg/s 1.50 1.50 0.20 0.20 1.30 1.30 1.30
H2 kg/s 0.03
O2 kg/s 0.39 1.13 0.20 0.20 0.38 1.43 1.43 16.78 6.39 0.36 0.05
SO2 kg/s 0.08 0.08
N2 kg/s 1.27 3.68 7.06 7.06 1.22 4.67 4.67 54.73 20.83 1.14
Ash kg/s 0.11 0.11 0.01 0.01 0.10 0.10 0.10 0.08
C kg/s 0.44
S kg/s
H2O kg/s 0.24 0.24 0.07
C2 S kg/s 7.32 7.32 0.81 0.81 5.33 5.33 5.33
C3 A kg/s 0.52 0.52 0.52 0.52 3.44 3.44 3.44
C3 S kg/s 2.85 2.85 2.85 2.85 18.88 18.88 18.88
C4AF kg/s 0.39 0.39 0.39 0.39 2.59 2.59 2.59
Fe2O3 kg/s 0.85 0.85
Al2O3 kg/s 1.84 1.84
SiO2 kg/s 4.65 4.65
AS4H kg/s
AS2H2 kg/s
2FeO⋅OH kg/s
CaO kg/s 16.16 16.16
Total Flow kg/s 1.66 4.81 36.76 36.76 14.98 14.98 1.60 31.72 31.72 6.10 6.10 71.52 31.72 27.22 1.50 0.68
Temperature °C 908 9 921 1109 1450 1025 9 1450 1370 1025 1400 9 60 279 9 9
Pressure atm >1 1 - - >1 >1 1 - - >1 >1 >1 - >1 >1 -
Total Stream MJ/h +5.80 -0.29 -1508 -1479 -225.2 -253.7 -0.09 -1259 -1270 +24.33 +35.17 -4.25 -1425 +25.82 -0.09 -7.7
Enthalpy
207
Stream Number 28 29 30 31 32 33 34 35 36 37 38 39 40

Stream Name Pet- Dry Dry Pet- B/F Exit Hot Air to Steam Purge CO2 depleted CaO to CaCO3 to CO2 rich Oxygen Make- Pet-
coke Coal Coke Gas Cycle gas Carbonator Calciner gas up coke
Component
CO2 kg/s 0.09 0.77 33.60 29.33
CaCO3 kg/s 0.07 15.92
CaSO4 kg/s 0.66 0.89 0.89
H2 kg/s 0.03 0.03 0.03 0.15
O2 kg/s 0.01 0.05 0.01 0.06 8.61 0.16 1.14 12.48 0.08
SO2 kg/s 0.01
N2 kg/s 0.01 0.01 1.50 28.10 12.02 0.13 0.05
Ash kg/s 0.08 0.01 0.01 0.01 0.01
C kg/s 0.65 0.44 0.65 3.74
S kg/s 0.04 0.04 0.16
H2O kg/s 0.01 0.19 1.02 3.99 2.55 0.08
C2 S kg/s
C3 A kg/s
C3 S kg/s
C4AF kg/s
Fe2O3 kg/s
Al2O3 kg/s
SiO2 kg/s 0.03
AS4H kg/s 0.04
AS2H2 kg/s 0.01
2FeO⋅OH kg/s
CaO kg/s 16.16 21.60 12.68
Total Flow kg/s 0.76 0.77 0.74 1.85 36.71 16.83 13.98 22.50 29.50 38.87 12.53 31.88 4.29
Temperature °C 9 110 110 110 646 930 650 930 650 930 9 9 9
Pressure atm - - - >1 >1 - >1 - - >1 >1 - -
Total Stream MJ/h +0.33 -10.98 +1.41 -11.75 +88.10 -634 -37.78 -847.3 -1176 -1122 -0.67 -1422 -0.23
Enthalpy
208
Figure A3 Schematic diagram of the steam cycle design (The below equation has been used to estimate the lumped conversion factor.
The adiabatic efficiencies in HP, IP and LP turbines were assumed to be 86%, 86% and 95%, respectively (Ahn et al., 2013) and that for the pump was
taken as 75%).

𝑊𝑖𝑑𝑒𝑎𝑙,𝐻𝑃 .𝜂𝑎𝑑,𝐻𝑃 +𝑊𝑖𝑑𝑒𝑎𝑙,𝐼𝑃 .𝜂𝑎𝑑,𝐼𝑃 +𝑊𝑖𝑑𝑒𝑎𝑙,𝐿𝑃 .𝜂𝑎𝑑,𝐿𝑃 −𝑊𝑖𝑑𝑒𝑎𝑙,𝑝𝑢𝑚𝑝 .𝜂𝑎𝑑,𝑝𝑢𝑚𝑝


𝜂= (Eq. A1)
𝑄𝑡𝑜𝑡𝑎𝑙
209
Table A3 Stream compositions for the preliminary steam cycle design.

Stream Molar Flow Temperature Pressure Total


Number [kmol/h] [°C] [bar] Stream
Enthalpy
[GJ/h]
1 1 565 167 -0.2244
2 1 347.3 38 -0.2312
3 1 565 38 -0.2221
4 1 469.1 20 -0.2256
5 1 28.98 0.04 -0.2460
6 1 25 0.04 -0.2850
7 1 25 0.04 -0.2850
8 1 26.69 167 -0.2846
Stream Power
(MWe)
Q-100 1.894e-3
Q-101 9.76e-4
Q-102 5.646e-3
Q-103 1.118e-4
Power
(MWth)
Q-104 1.674e-2
Q-105 2.511e-3
Condenser 1.085e-2
Energy
210
Appendix B Carbonator Design and Customization of
UniSim Design

B.1 Carbonator Design

In the simple model, the maximum average carbonation degree of sorbent in the
solid population is defined as Eq. B1 (Abanades, 2002) where Xmax,N is the
maximum carbonation degree after N cycles of carbonation/calcination in Eq. B2
(Grasa and Abanades, 2006) and rN is the mass fraction of particles calculated from
mass balance in Eq. B3 (Abanades, 2002). The CO2 capture efficiency in the
carbonator (ECO2) is defined as Eq. B4.

𝑋𝑚𝑎𝑥,𝑎𝑣𝑒 = ∑𝑁=∞
𝑁=1 𝑟𝑁 𝑋𝑚𝑎𝑥,𝑁 (B1)
1
𝑋𝑚𝑎𝑥,𝑁 = + 𝑋𝑟 (B2)
1⁄(1−𝑋𝑟 )+𝑘𝑁

𝐹0 𝐹𝑅𝑁−1
𝑟𝑁 = (𝐹0 +𝐹𝑅 )𝑁
(B3)

𝐹𝑅 𝑋𝑚𝑎𝑥,𝑎𝑣𝑒
𝐸𝐶𝑂2 = (B4)
𝐹𝐶𝑂2

Xr and k are constants specific to the type of limestone. For non-sulfated Piaseck
limestone the values of the constants are Xr=0.0969 and k=0.66 (Grasa et al., 2008b;
Romano, 2012). The capture efficiency calculated from this model is shown in
Figure B1. The maximum carbonation efficiency is also limited by the equilibrium of
CO2 over CaO as shown in Eq. B5 (Garcia-Labiano et al., 2002). The equilibrium at
650°C is shown as a horizontal dotted line in Figure B1.

20474
𝑃𝐶𝑂2,𝑒𝑞 = 4.137 × 1012 exp (− ) (B5)
𝑇

As it can be noticed from Figure B1, various sets of F0/FCO2 and FR/FCO2 can
provide capture efficiency of 90% in the carbonator according to Eqs. B1 to B4.
However, an advanced carbonator model is necessary in order to deal with the
hydrodynamics of a fluidized bed system.

211
1

Eequil

0.8

0.6
ECO2

FR/FCO₂=1
FR/FCO₂=1.2
0.4
FR/FCO₂=2
FR/FCO₂=3
FR/FCO₂=4

0.2

0
0 1 2 3 4 5
F0/FCO2

Figure B1 Graphical explanation of Eq. B5 at 650°C (k=0.66 and Xr=0.0969).

Detailed carbonator designs have been investigated including the design of a BFB
(Abanades et al., 2004) and a fast fluidized bed (Lasheras et al., 2011; Romano,
2012). Therefore, in this study, the detailed CFB model for fast fluidization has been
employed based on the approach proposed in the literature (Rampinelli, 2010;
Romano, 2012) and named rigorous model. This model is briefly described here but
further details can be found in the literature (Romano, 2012).

In the rigorous model, the flow regime needs to be determined before the
prediction of the behaviour of gas-solid reaction. Kunii and Levenspiel (1991)
constructed Figure B2 for the whole range of gas-solid contact regimes while the
letters A, B, C and D refer to Geldart classification of the solids.

212
Figure B2 General flow regime diagram for a gas-solid system (Kunii and
Levenspiel, 1991).

The dimensionless particle size 𝑑𝑝∗ and the dimensionless gas velocity 𝑢∗ can be
calculated to map the constructed chart according to the following equations:

−1
∗ 18 0.591
𝑢 =� ∗ )2 + ∗ )0.5 � for spherical particles (B6)
(𝑑𝑝 (𝑑𝑝

1/3
𝜌𝑔 �𝜌𝑠 −𝜌𝑔 �𝑔
𝑑𝑝∗ = 𝑑𝑝 � � (B7)
𝜇2

where dp is the particle size of a sorbent and μ is the viscosity. ρg and ρs refer to gas
and solid densities, respectively.

The solid distribution part of the rigorous model has been based on the Kunii-
Levenspiel model for CFBs (Kunii and Levenspiel, 1991), and the reactor is divided
into two sections, a lower dense region and an upper lean region. To calculate the
values of Hl (height of the upper lean region) and εs,e (exiting solid fraction), the
particle distribution equations (Eqs. B8 and B9) need to be solved simultaneously.

213
𝜀𝑠,𝑒 = 𝜀𝑠∗ + (𝜀𝑠,𝑑 − 𝜀𝑠∗ )𝑒 −𝑎𝐻𝑙 (B8)

𝑊𝑠 �𝜀𝑠,𝑑 −𝜀𝑠,𝑒 �
= + 𝐻𝑡 𝜀𝑠,𝑑 − 𝐻𝑙 �𝜀𝑠,𝑑 − 𝜀𝑠∗ � (B9)
𝐴𝑡 𝜌𝑠 𝑎

where 𝜀𝑠∗ is the asymptotic solid volumetric fraction and is assumed to be 0.01 for a
fast fluidized bed. 𝜀𝑠,𝑑 is the volume fraction of solids in the lower dense region and
is set equal to 0.15. ‘a’ is the decay constant of solid concentration in the lean region,
and to estimate this constant, a value of 3 s-1 for a.u0 is taken from the values ranging
from 2 to 7 s-1 in Kunii and Levenspiel (1991). Ws is the solid inventory in the
carbonator and determined by setting the pressure drop in the carbonator to 0.1 bar.

The solution of the gas phase material balance in the CFB which can be
rearranged with the first order kinetic law for the carbonation reaction (Eq. 2.5 in
Chapter 2) leads to two final equations, B10 and B11 which give CO2 concentrations
at the top of the dense region (CCO2,d) and at the reactor exit (CCO2,out), respectively.

𝐶𝐶𝑂2,𝑑 =

𝐶𝐶𝑂2,𝑒𝑞 + (𝐶𝐶𝑂2,𝑖𝑛 − 𝐶𝐶𝑂2,𝑒𝑞 )𝑒 −�𝜉𝜀𝑠,𝑐 𝛿𝑘𝑟𝑖,𝑎𝑣𝑒+1/((1⁄𝛿𝐾𝑐𝑤)+1⁄(𝜉𝜀𝑠,𝑤(1−𝛿)𝑘𝑟𝑖,𝑎𝑣𝑒))�𝐻𝑑 /𝑢0 (B10)

𝐶𝐶𝑂2,𝑜𝑢𝑡 = 𝐶𝐶𝑂2,𝑒𝑞 + (𝐶𝐶𝑂2,𝑑 − 𝐶𝐶𝑂2,𝑒𝑞 )𝑒 −�𝑘𝑟𝑖,𝑎𝑣𝑒⁄𝑢0�(𝑥+𝑦) (B11)

1−𝜂𝑠𝑑
𝑥 = 𝜉𝜀𝑠∗ �𝐻𝑙 − (1 − 𝑒 −𝑏𝐻𝑙 )� (B12)
𝑏

1−𝑒 −𝑎𝐻𝑙 1−𝜂𝑠𝑑


𝑦 = 𝜉(𝜀𝑠,𝑑 − 𝜀𝑠∗ ) � − (1 − 𝑒 −(𝑎+𝑏)𝐻𝑙 )� (B13)
𝑎 𝑎+𝑏

The average kinetic constant (kri,ave) and volume fraction of potentially active
solids (ξ) should be known initially to solve Eqs. B10 and B11. Therefore, as another
approach, the considerations of solid composition in the carbonator, probability
density function (ft), i.e. the fraction of particles with certain residence times, in Eqs.
B14-16 and the distribution of the particle based on the number of carbonation-
calcination cycles as given in Eq. B3 allow the calculation of the average carbonation
level, Xave, (Eq. B17) as well as kri,ave (Eq. B18) and ξ.

1
𝑓𝑡 = 𝑒 −(𝑡⁄𝜏) (B14)
𝜏

214
𝑛𝑠,𝑎
𝜏= (B15)
𝐹𝑅

𝑊𝑠
𝑛𝑠,𝑎 = (1 − 𝑥𝑎𝑠ℎ − 𝑥𝐶𝑎𝑆𝑂4 ) (B16)
𝑀𝑠

𝑡 ∞
𝑋𝑎𝑣𝑒 = ∑+∞
𝑁=1 𝑟𝑁 �∫0
𝑙𝑖𝑚 ∗
𝑓𝑡 𝑋(𝑡, 𝑁, 𝐶𝐶𝑂2
)𝑑𝑡 + ∫𝑡 𝑓𝑡 𝑋𝑚𝑎𝑥,𝑁 𝑑𝑡� (B17)
𝑙𝑖𝑚

𝜌 𝑡
𝑘𝑟𝑖 , 𝑎𝑣𝑒 = 𝑠,𝑎 ∑+∞ ∗
𝑟 ∫ 𝑙𝑖𝑚 𝑓𝑡 𝑘𝑠 𝑆𝑁 (1 − 𝑋�𝑡, 𝑁, 𝐶𝐶𝑂 �)2/3 𝑑𝑡 (B18)
𝑀𝑠,𝑎 𝑁=1 𝑁 0 2

It should be noted that, the effect of sulphation on the maximum carbonation


degree in the rigorous model was considered by adjusting the k and Xr constants (Eq.
B2) corresponding to the sulphation level of Piaseck limestone (Grasa et al., 2008b;
Romano, 2012). The given equations are only valid up to a sulphation level of 1%.

∆𝑋𝐶𝑎𝑆𝑂4 = 𝐹𝑆 /(𝐹𝑅 + 𝐹0 ) (B19)

𝑘 = 0.026 × (∆𝑋𝐶𝑎𝑆𝑂4 × 100)2 + 0.219 × �∆𝑋𝐶𝑎𝑆𝑂4 � + 0.660 0 ≤ ∆𝑋𝐶𝑎𝑆𝑂4 ≤ 0.01 (B20)

𝑋𝑟 = (−0.1118 × ∆𝑋𝐶𝑎𝑆𝑂4 × 100) + 0.0969 0 ≤ ∆𝑋𝐶𝑎𝑆𝑂4 ≤ 0.005 (B21)

𝑋𝑟 = (−0.0298 × ∆𝑋𝐶𝑎𝑆𝑂4 × 100) + 0.0559 0.005 < ∆𝑋𝐶𝑎𝑆𝑂4 ≤ 0.01 (B22)

The capture efficiency can be calculated by two different ways, Eq. B23 and Eq.
B24.

𝐹𝑅 𝑋𝑎𝑣𝑒
𝐸′𝐶𝑂2 = (B23)
𝐹𝐶𝑂2

𝐹𝐶𝑂2 −𝑉𝑔,𝑜𝑢𝑡 𝐶𝐶𝑂2,𝑜𝑢𝑡 𝑉𝑔,𝑖𝑛 𝐶𝐶𝑂2,𝑖𝑛 −𝑉𝑔,𝑜𝑢𝑡 𝐶𝐶𝑂2,𝑜𝑢𝑡


𝐸′′𝐶𝑂2 = = (B24)
𝐹𝐶𝑂2 𝑉𝑔,𝑖𝑛 𝐶𝐶𝑂2,𝑖𝑛

The values for the variables required for the carbonator calculation are imported
from the cement plant simulation via a COM interface (FCO2, F0, FR, Fash, FS, u0, dp,
T, Ms, p, μ, CCO2,in, Vg,in, Ws, Ht, xCaSO4, xash). The area of the reactor (At), average
solid density (ρs) and average molar mass values (Ms) are initially calculated. By
assuming a CO2 concentration inside the carbonator (CCO2*), Xave is obtained using
Eqs. B14 to B17 and the first capture efficiency, E’CO2 is calculated using Eq. B23.

215
The length of dense and lean regions is determined by Eqs. B8 to B9, given the
total height of carbonator, Ht. The average kinetic constant of the carbonation
reaction, kri,ave, can be calculated using Eq. B18 with the same CCO2* used in the
E’CO2 calculation. Using the kri,ave, the CO2 concentration at the outlet is calculated by
Eqs. B10 to B13. Finally the second capture efficiency, E″CO2 can be obtained using
Eq. B24. An iterative calculation is applied to obtain same capture efficiency from
the two different methods. Once the same capture efficiency is reached, it is possible
to calculate a new At using the average value of inlet and outlet flowrates and a new
Ms based on the Xave. Based on the new At and Ms, the new CCO2* needs to be
calculated to give same capture efficiency, ie. another iterative loop is set to obtain
At and Ms outside the iterative loop for CCO2*.

B.2 Customization of UniSim Design

The behaviour of a user defined operation in the UniSim Design can be defined by
compiling a Visual Basic code. Two different ways of implementing an external
program into UniSim environment are proposed. These methods include direct
integration of a Matlab code or an executable file as a solver. The basic user defined
operation appears as shown in Figure B3. The design tab of this unit contains three
main sections: connections, code and variables among which the last two are crucial
for the customization. A Visual Basic code is written inside the code section while
the external user defined parameters, i.e. height of a reactor and solid particle
diameter for the carbonator design, can be identified in the variables section. The
code environment comprises of three sub-titles as follows.

B.2.1 Sub Initialize

It is the section to activate the inlet and outlet streams and to set up the unit
operation. The below command allows the activation of the first feed stream that is
named ‘Gas Feed’. Similar commands can be repeated for the activation of all inlet
and outlet streams.

- ActiveObject.Feeds1Name = "Gas Feed"

216
Figure B3 The view of a user defined operation in UniSim Design.

B.2.2 Sub Execute

It is the main section to transfer of the stream properties into the unit operation. In
addition, a user defined variable can be created as an external input. The model
equations are compiled in this section while any changes on the stream properties or
the external inputs trigger the solver. The following command is employed to create
a user defined variable called “Pressure”, and the term ‘uctPressure’ assigns a unit
selection for this variable.

- ActiveObject.CreateUserVariable("Prs", "Pressure", uvtReal, uctPressure,


Scalar)

A crutial step, after compiling model equations, is the transfer of variables from
UniSim environment to an external program. Two different procedures are followed
for this purpose. First, if the external program is Matlab, the following commands: (i)
activate Matlab solver that runs in the background, (ii) add a path to the folder
referred as “folder” under C:\, and (iii) assign “MF1” value in the user defined
operation as “Massgas” in the Matlab environment. In the final line, the Matlab file
named “sample” is executed.

- Dim MatLab As Object


- Dim Result As String
- Set MatLab = CreateObject("Matlab.Application")

217
- Result = MatLab.Execute("cd c:\folder")
- Call MatLab.PutWorkspaceData("Massgas", "base", MF1)
- Result = MatLab.Execute("sample")

The assigned “Massgas” variable in the Matlab Workspace can be referred as


“Massgas_M” by using the below command if it is needed.

- Massgas_M = evalin('base','Massgas')

After Matlab solver terminates, the outputs can be sent to UniSim environment by
using “assignin” command. In the following example, the variable “M_result” from
Matlab is assigned as “U_result” in UniSim Design.

- assignin('base','U_result',M_result)

If an external model is written in any other programming languages in the form


of an executable file, the following procedure is applicable. This option is more
challenging since transferred data needs to be stored in input and output files. It is
also possible to adapt an executable file from a Matlab code if it is converted to an
executable by using Matlab Compiler.

(1) Any variables from a unit operation can be transferred into an input file by
using the following command. Here, the input file, “file.txt”, is initially cached.
Then, the term “ref” in the input file is replaced by a variable, “U_value”.
- sname = "C:\folder\file.txt"
- lOpenFile = FreeFile
- Open sname For Input As lOpenFile
- cached = Input(LOF(lOpenFile), lOpenFile)
- cached = Replace(cached, "ref", U_value)
(2) The executable solver (“solver.exe”) that requires an input file (“file.txt” in this
case) can be run by using the following:
- Dim wsh As Object
- Set wsh = CreateObject("WScript.Shell")
- Dim waitOnReturn As Boolean
- waitOnReturn = True

218
- Dim windowStyle As Integer
- windowStyle = 1
- wsh.Run("C:\folder\solver.exe C:\folder\file",windowStyle, waitOnReturn)
UniSim solver waits until the external solver stops running.
(3) In the final stage, any outputs of the external solver in the form of a text file
can be read by UniSim solver. In the final command line, the variable ‘Value’
is set equal to the second variable of the output (srtCDRackk).
- sFileName = "c:\folder\output.txt"
- lOpenFile = FreeFile
- Open sFileName For Input As lOpenFile
- sFileText = Input(LOF(lOpenFile), lOpenFile)
- strCDRackk = Split(sFileText)
- Close lOpenFile
- Value=strCDRackk(2)

The returned variables from an external model using the above implementation
methods can be then employed in the unit operation or transferred to the outlet
streams. For each outlet stream; temperature, pressure, mass or molar flow rate and
composition values have to be provided for activation. The following command is to
set the temperature of ‘SolidOut’ stream to the value of ‘Mtemp’.

- SolidOut.Temperature.Calculate(Mtemp)

B.2.3 Sub StatusQuery

This section is to update of status information. Any error and warning messages
for missing streams, connections or external variables can be assessed. The error
messages not only appear in UniSim status bar but also are illustrated at the bottom
of the user defined operation. The final form of the implementation of the Ca-looping
process using an external carbonator model and the developed interface are presented
in Figure B4. The proposed integration allows rapid data transfer whenever any
model inputs are changed, and the customised error messages let the user to
understand the source of an error.

219
(a)

(b)

Figure B4 a) The view of Ca-looping process in the simulation environment, and b)


the interface of the carbonator in UniSim.

220
Appendix C

A sample economic analysis spreadsheet for the Ca-looping

process

221
Table C1 Economic analysis spreadsheet for the Ca-looping process at 1.65 F0/FCO2.
2013 2014 2015 2016 2017 2018 2019 2020 2021 2022 2023 2024 2025 2026 2027 2028 2029 2030 2031 2032 2033 2034 2035 2036 2037 2038 2039 2040 2041
Year 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26
Load Factor 60 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90
Equivalent yearly hours 5256 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884 7884
Expenditure Factor 20 45 35

Operating Costs
Fuel 9.47 14.41 14.63 14.85 15.07 15.29 15.52 15.76 15.99 16.23 16.48 16.72 16.97 17.23 17.49 17.75 18.02 18.29 18.56 18.84 19.12 19.41 19.70 20.00 20.30 0
Power 0.24 0.37 0.38 0.38 0.39 0.40 0.40 0.41 0.41 0.42 0.43 0.43 0.44 0.45 0.45 0.46 0.47 0.47 0.48 0.49 0.49 0.50 0.51 0.52 0.53 0
Raw materials + Cooling water 4.48 6.82 6.92 7.02 7.13 7.23 7.34 7.45 7.56 7.68 7.79 7.91 8.03 8.15 8.27 8.39 8.52 8.65 8.78 8.91 9.04 9.18 9.32 9.46 9.60 0

ETS 0.20 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0.30 0

Maintenance 11.23 17.09 17.35 17.61 17.87 18.14 18.41 18.69 18.97 19.25 19.54 19.84 20.13 20.44 20.74 21.05 21.37 21.69 22.02 22.35 22.68 23.02 23.37 23.72 24.07 0
Labour & Supervision 4.38 6.66 6.76 6.86 6.97 7.07 7.18 7.29 7.40 7.51 7.62 7.73 7.85 7.97 8.09 8.21 8.33 8.46 8.58 8.71 8.84 8.97 9.11 9.25 9.38 0
Administration, Local Rates&Insurance 6.93 10.55 10.70 10.86 11.03 11.19 11.36 11.53 11.70 11.88 12.06 12.24 12.42 12.61 12.80 12.99 13.18 13.38 13.58 13.79 13.99 14.20 14.42 14.63 14.85 0
Fixed Capital Expenditures 91.79 206.52 160.62

Total cash flow (yearly) 91.79 206.52 197.54 56.20 57.04 57.89 58.76 59.64 60.53 61.43 62.35 63.28 64.22 65.18 66.15 67.14 68.14 69.16 70.19 71.24 72.31 73.39 74.48 75.59 76.72 77.87 79.03 0
Total cash flow (cumulated) 91.79 298.30 495.85 552.05 609.09 666.99 725.75 785.38 845.91 907.34 969.68 1032.96 1097.18 1162.36 1228.51 1295.65 1363.79 1432.95 1503.14 1574.38 1646.69 1720.08 1794.56 1870.15 1946.87 2024.74 2103.78 2103.78

Cement production 0 0 0 0.60 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 0.90 22.20
Sum
Discount rate 0.1
Discount factor 1 0.91 0.83 0.75 0.68 0.62 0.56 0.51 0.47 0.42 0.39 0.35 0.32 0.29 0.26 0.24 0.22 0.20 0.18 0.16 0.15 0.14 0.12 0.11 0.10 0.09 0.08 0.08

Total cash flow 83.44 170.68 148.42 38.39 35.42 32.68 30.15 27.82 25.67 23.68 21.85 20.16 18.60 17.16 15.84 14.61 13.48 12.44 11.48 10.59 9.77 9.02 8.32 7.67 7.08 6.53 6.03 826.98
Output 0.45 0.61 0.56 0.51 0.46 0.42 0.38 0.35 0.32 0.29 0.26 0.24 0.22 0.20 0.18 0.16 0.15 0.13 0.12 0.11 0.10 0.09 0.08 0.08 0.07 6.53

Operating sum 10.66 14.75 13.61 12.56 11.59 10.69 9.87 9.11 8.40 7.75 7.15 6.60 6.09 5.62 5.19 4.79 4.42 4.07 3.76 3.47 3.20 2.95 2.73 2.51 2.32
Fixed sum 16.93 23.43 21.62 19.95 18.41 16.98 15.67 14.46 13.34 12.31 11.36 10.48 9.67 8.93 8.24 7.60 7.01 6.47 5.97 5.51 5.08 4.69 4.33 3.99 3.68
Capital 83.44 170.68 120.68
ETS 0.15 0.21 0.19 0.17 0.16 0.14 0.13 0.12 0.11 0.10 0.09 0.08 0.07 0.07 0.06 0.05 0.05 0.05 0.04 0.04 0.03 0.03 0.03 0.03 0.02

Summary
Operating (€/ton cement) 26.64
Fixed (€/ton cement) 42.31 Cost of base 97.33 (€/t cement)
Capital (€/ton cement) 57.43 CO2 avoided 41.11 €/ton CO2 avoided
ETS (€/ton cement) 0.34
Sum 126.72
222
Appendix D

Stream properties for the steam cycle of biomass-air-fired

power plant, and CO2 purification and compression unit,

and the effect of fuel moisture on gross capacity of the

biomass-air-fired plant described in Chapter 5

223
Table D1 The steam cycle stream properties for the biomass-air-fired plant presented in Figure 5-2(b).

Stream No 1 2 3 4 5 6 7 8 9 10
T (°C) 565.6 565.6 565.6 565.6 565.6 365.9 365.9 365.9 565.6 496.9
P (bar) 38.31 166.5 166.5 166.5 38.31 42.78 42.78 42.78 166.5 24.16
m (ton/h) 368.5 401.6 0.11 0.41 368.9 0.40 0.80 396.8 3.1 18.60
Stream No 11 12 13 14 15 16 17 18 19 20
T (°C) 395.2 395.2 395.2 395.2 110.7 276.9 229.6 72.0 38.73 46.40
P (bar) 11.56 11.56 11.56 11.56 0.81 4.42 2.87 0.34 0.069 0.103
m (ton/h) 28.31 325.1 308.9 16.26 11.12 8.65 20.08 10.5 258.5 16.26
Stream No 21 22 23 24 25 26 27 28 29 30
T (°C) 395.7 69.67 38.39 38.68 56.74 78.19 116.3 131.7 130.9 93.61
P (bar) 42.78 0.34 0.069 27.23 27.23 27.23 27.23 27.23 4.42 2.87
m (ton/h) 0.51 50.35 325.6 325.6 325.6 325.6 325.6 325.6 8.65 28.73
Stream No 31 32 33 34 35 36 37 38
T (°C) 72.44 184 188.3 218.1 251.5 223.4 193.8 365.9
P (bar) 0.81 10.98 214.5 214.5 214.5 42.78 24.16 42.78
m (ton/h) 39.84 401.6 401.6 401.6 401.6 28.26 46.86 368.5
224
25 160

20

Gross Power (MWe)


120
LHV (MJth/kg)
15
80
10

40
5

0 0
0 10 20 30 40 50 60 70

Moisture Content (%)


Figure D1 The effect of moisture content on the LHV (MJth/kg biomass) of the biomass and gross power generation capacity of the biomass-air-fired plant. Dashed
line corresponds to the gross power with the same mass of biomass on a dry basis. Continuous line corresponds to the gross power with the same mass of wet biomass.
225
Table D2 The stream properties and compositions for the cryogenic CO2 purification and compression unit (Figure 5-3) added to the biomass-oxy-fired plant.

Stream No 1 2 3 4 5 6 7 8 9 10
T (°C) 80 34.7 35 −20 −35 −35 −32 28.2 −35 33.2
P (bar) 1.013 1.013 24 24 24 24 80 80 24 54
Molar Flow Rate (kgmole/h) 4168.3 1832.1 2336.2 2336.2 2336.2 1748.4 1748.4 1748.4 587.8 587.8
Composition (mole %)
CO2 48.9 1.0 86.5 86.5 86.5 97.2 97.2 97.2 54.8 54.8
O2 3.4 - 6.1 6.1 6.1 1.3 1.3 1.3 20.5 20.5
N2 4.0 - 7.1 7.1 7.1 1.2 1.2 1.2 24.7 24.7
H2O 43.7 99.0 0.3 0.3 0.3 0.3 0.3 0.3 - -
Total Stream Enthalpy (GJth/h) −1235.0 −527.4 −798.6 −811.3 −828.5 −700.1 −699.6 −686.8 −128.5 −127.3
Stream No 11 12 13 14 15 16 17 18 19 20
T (°C) −18 −35 −35 −33.2 −35 8 25 15 26.3 43.4
P (bar) 54 54 54 80 54 54 1.12 80 80 150
Molar Flow Rate (kgmole/h) 587.8 587.8 225.0 225.0 362.8 362.8 362.8 225.0 1973.4 1973.4
Composition (mole %)
CO2 54.8 54.8 91.3 91.3 32.1 32.1 32.1 91.3 96.5 96.5
O2 20.5 20.5 4.5 4.5 30.4 30.4 30.4 4.5 1.7 1.7
N2 24.7 24.7 4.2 4.2 37.5 37.5 37.5 4.2 1.5 1.5
H2O - - - - - - - - 0.3 0.3
Total Stream Enthalpy (GJth/h) −129.3 −131.5 −84.4 −84.3 −47.1 −46.5 −45.9 −83.0 −769.9 −768.7
226
Appendix E

E.1 Publications

• Ozcan, D.C., Alonso, M., Ahn, H., Abanades, J.C., Brandani, S., 2014. Process and
Cost Analysis of a Biomass Power Plant with In-situ Calcium Looping CO2 Capture
Process. Industrial & Engineering Chemistry Research 53 (26), 10721-10733.
• Ozcan, D.C., Ahn. H., Brandani, S., 2013. Process Integration of a Ca-looping Process
with a Cement Plant for Carbon Capture. International Journal of Greenhouse Gas
Control 19, 530-540.
• Romano, M. C., Anantharaman, R., Arasto, A., Ozcan, D.C., Ahn, H., Dijkstra, J.W.,
Carbo, M., Boavida, D., 2013. Application of Advanced Technologies for CO2
Capture from Industrial Sources. Energy Procedia 37, 7176-7185.
• Romano, M.C., Martínez, I., Murillo, R., Arstad, B., Blom, R., Ozcan, D.C., Ahn, H.,
Brandani, S., 2013. Process Simulation of Ca-Looping Process: Review and
Guidelines. Energy Procedia 37, 142-150.
• Romano, M.C., Martínez, I., Murillo, R., Arstad, B., Blom, R., Ozcan, D.C., Ahn, H.,
Brandani, S., 2012. Guidelines for modelling and simulation of Ca-looping processes.
EERA CCS report.

E.2 Presentations

• Ozcan, D.C., 2014. Carbon Capture from Industrial Sources. Association of British
and Turkish Academics (ABTA) Doctoral Researcher Awards Ceremony, May 25,
London, UK.
• Ozcan, D.C., Bocciardo, D., Ferrari, M.C., Ahn, H., Kierzkowska, A.M., Muller,
C.R., Brandani, S., 2013. Comparison of Various Carbon Capture Technologies to
Reduce CO2 emissions From a Cement Plant. AIChE 2013 Annual Meeting,
November 3-8, San Francisco, USA.
• Ozcan, D.C., 2013. Selection of an appropriate carbon capture technology for the
reduction of CO2 emissions from a cement plant. 25th Annual Honeywell Users Group
EMEA Conference, November 4-7, Nice, France.

227
• Ozcan, D.C., Ahn, H., Kierzkowska, A.M., Muller, C.R., Brandani, S., 2013. Process
Integration of Chemical Looping Combustion into Calcium Looping for CO2 Capture
from a Cement Plant. 5th IEAGHG HTSLCN Meeting, September 2-3, Cambridge,
UK.
• Ozcan, D.C., Ahn, H., Brandani, S., 2012. Techno-Economic Study of Ca-Looping
Processes for Carbon Capture from Cement Plants. AIChE 2012 Annual Meeting,
October 28 – November 2, Pittsburgh, USA.
• Ozcan, D.C., Ahn, H., Brandani, S., 2012. Fully integrated simulation of a cement
plant with a carbon capture Ca-looping process. 4th IEAGHG HTSLCN Meeting,
August 19-21, Beijing, China.
• Ozcan, D.C., Ahn, H., Brandani, S., 2011. Design of a Ca-looping Process for Carbon
Capture in a Cement Plant. CCS – Development of Improved Technologies for
Chemical Looping Combustion & Carbonate Looping Meeting, November, Zabrze,
Poland.
• Ozcan, D.C., Ahn, H., Brandani, S., 2011. Process Integration of a Ca-looping process
with a cement manufacturing plant. Joint UKCCSRC and CO2CHEM Meeting, July,
Nottingham, UK.

E.3 Honours & Awards

• Honorable Mention Award from Association of British and Turkish Academics


(ABTA) in 2014.
• Winner of the Honeywell UniSim® Design Challenge for the Europe, Middle East,
Africa (EMEA) region in 2013.

228

View publication stats

You might also like